pyrrolizidine alkaloids: chemical basis of toxicity...2 results of separation of pyrrolizidine...

161
Pyrrolizidine alkaloids: Chemical basis of toxicity Item Type text; Dissertation-Reproduction (electronic) Authors Cooper, Roland Arthur, 1963- Publisher The University of Arizona. Rights Copyright © is held by the author. Digital access to this material is made possible by the University Libraries, University of Arizona. Further transmission, reproduction or presentation (such as public display or performance) of protected items is prohibited except with permission of the author. Download date 18/04/2021 09:18:33 Link to Item http://hdl.handle.net/10150/290581

Upload: others

Post on 01-Nov-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

Pyrrolizidine alkaloids: Chemical basis of toxicity

Item Type text; Dissertation-Reproduction (electronic)

Authors Cooper, Roland Arthur, 1963-

Publisher The University of Arizona.

Rights Copyright © is held by the author. Digital access to this materialis made possible by the University Libraries, University of Arizona.Further transmission, reproduction or presentation (such aspublic display or performance) of protected items is prohibitedexcept with permission of the author.

Download date 18/04/2021 09:18:33

Link to Item http://hdl.handle.net/10150/290581

Page 2: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

INFORMATION TO USERS

This manuscript has been reproduced from the microfikn master. UMI

films the text directly from the original or copy submitted. Thus, some

thesis and dissertation copies are in typewriter face, while others may be

from any type of computer printer.

The quality of this reproduction is dependent upon the quality of the

copy submitted. Broken or indistinct print, colored or poor quality

illustrations and photographs, print bleedthrough, substandard margins,

and improper alignment can adversely affect reproduction.

In the unlikely event that the author did not send UMI a complete

manuscript and there are missing pages, these will be noted. Also, if

unauthorized copyright material had to be removed, a note will indicate

the deletion.

Oversize materials (e.g., maps, drawings, charts) are reproduced by

sectioning the original, beginning at the upper left-hand comer and

continuing from left to right in equal sections with small overlaps. Each

original is also photographed in one exposure and is included in reduced

form at the back of the book.

Photographs included in the original manuscript have been reproduced

xerographically in this copy. Higher quality 6" x 9" black and white

photographic prints are available for any photographs or illustrations

appearing in this copy for an additional charge. Contact UMI directly to

order.

UMI A Bell & Howell Information Company

300 North Zeeb Road, Ann Aibor MI 48106-1346 USA 313/761-4700 800/521-0600

Page 3: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet
Page 4: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

PYRROLIZIDINE ALKALOIDS: CHEMICAL BASIS OF TOXICITY

by

Roland Arthur Cooper

A Dissertation Submitted to the Faculty of the

COMMITTEE ON PHARMACOLOGY AND TOXICOLOGY (GRADUATE)

In Partial Fulfillment of the Requirements For the Degree of

DOCTOR OF PHILOSOPHY

In the Graduate College

THE UNTVERSITY OF ARIZONA

1 9 9 6

Page 5: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

UMI Nvunber: 9706164

UMI Microform 9706164 Copyright 1996, by UMI Company. All rights reserved.

This microform edition is protected against unauthorized copying under Title 17, United States Code.

UMI 300 North Zeeh Road Ann Arbor, MI 48103

Page 6: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

2

THE UNIVERSITY OF ARIZONA ® GRADUATE COLLEGE

As members of the Final Examination Committee, we certify that we have

read the dissertation prepared by Roland Arthur Cooper

entitled Pvrrolizidine Alkaloids: Giemical Basis of Toxicity

and recommend that it be accepted as fulfilling the dissertation

requirement for the Degree of Doctor of Philosophy

>-1 h -> JfG

~J ~A Date

Date

Date

Final approval and acceptance of this dissertation is contingent upon

the candidate's submission of the final copy of the dissertation to the

Graduate College.

I hereby certify that I have read this dissertation prepared under my

direction and recommend that it be accepted as fulfilling the dissertation

requii?ement.

Dissertation /Director

/ ^ / - 6 Date '

Page 7: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

3

STATEMENT BY AUTHOR

This dissertation has been submitted in partial fulfillment of requirements for an advanced degree at The University of Arizona and is deposited in the University Library to be made available to borrowers under rules of the Library.

Brief quotations from this dissertation are allowable without special permission, provided diat accurate acknowledgment of source is made. Requests for permission for extended quotation from or reproduction of this manuscript in whole or in part may be granted by the head of the major department or the Dean of the Graduate College when in his or her judgement the proposed use of the material is in die interests of scholarship. In all other instances, however, permission must be obtained by the author.

SIGNED: i

Page 8: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

4

ACKNOWLEDGMENTS

This thesis is dedicated to my family, who have provided constant encouragement,

patience, guidance and love throughout my college years.

A special thanks goes to my advisor, Dr Ryan Huxtable, for his support,

friendship, inspiration, some fine dinners and a wonderful trip to Brazil. Also to the rest of

my graduate committee. Dr. Hugh Laird, Dr. Klaus Brendel, Dr. Dan Liebler, Dr. Jim

Halpert, and Dr. Ray Nagle.

I thank many individuals for their assistance with various matters: Dr. Glenn

Sipes, Dr. Frank Porreca, Dr. Ron Lukas, Richard Felger, Carla Hayes, Jane Ageton, and

Sandi Sledge. I also thank Dr. C.C. Yan for the valuable data from the isolated perfused

liver.

I have made many friends while at the University of Arizona who have helped

make my stay in Tucson a lot of fun, from the parties to the great fishing trips in the White

Mountains: Gina Escobar, Steve Barr, Brent Barber, Ed Bilsky, Mike Nichols, Steve

Stratton, Todd Vanderah, Di Bian, Carla Beckham, John Murphy, Rion Bowers, Pierre-

Louis Lieu, Bob Horvath, Sandy Wegert, C.J. Kovelowski, just to mention a few.

Page 9: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

5

TABLE OF CONTENTS

LIST OF FIGURES 9

LIST OF TABLES 11

ABSTRACT 12

INTRODUCTION 14

Plant Sources of PAs 15

PA Exposure 15

PA Isolation, Separation and Puriflcation 18

Extraction from plant material 18

Spectrophotometric determination of PAs 18

Preparative chromatographic separation of PAs 19

High-speed counter-current chromatography 19

PA Structure 20

Chemical Properties of DHAs 24

PA Metabolism 25

N-Oxidation 26 Hydrolysis 26 Dehydrogenation to pyrroles 27

Toxic Effects of PAs 28

Hepatotoxicity 29 Pneumotoxicity 33 Neurotoxicity 34 Carcinogenicity and anti-tumor activity 35 Mutagenicity and cytotoxicity 35

Metabolites Responsible for Extrahepatic Toxicity 37

DHAs 37 Sulfur conjugated pyrroles: 7-GSDHP and 7-CYSDHP 38 DHP 39

Effect of PA Structure on Metabolism and Toxicity 40

Lipophilicity 40 Ester hydrolysis 40 Bioactivation to DHAs versus N-oxidation 41

Page 10: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

6

TABLE OF CONTENTS - Cont.

DHA reactivity 42

Statement of the Research Problem and Hypothesis 43

METHODS 46

PA Isolation, Separation and IdentiHcation 46

Plant material 46 Alkaloid isolation 46 High-speed counter-current chromatography 47 Thin-layer chromatography 48 Gas chromatography electron impact-mass spectrometry, method 1 48 Gas chromatography electron impact-mass spectrometry, method II 49 Flow-injection atmospheric pressure chemical ionization mass spectrometry 49

Preparation of PA Metabolites and Derivatives 49

DMAs 49 7-CYSDHP (7-cysteinyl-6,7-dihydro-I-hydroxymethyl-5H-pyrrolizine) 50 7-GSDHP (7-glutathionyl-6,7-dihydro-I-hydroxymethyl-5H-pyrrolizine) 50 DHP ((±)-6,7-dihydro- 7-hydroxy-1 -hydroxymethyl-SH-pyrrolizine) 50 9-Ethoxy DHP (7-hydroxy-9-ethoxy-6,7-dihydro-5H-pyrrolizine) 51

7-Ethoxy DHP ((±}-7-ethoxy-I-hydroxymethyl-6,7-dihydro-5H-pyrrolizine).. .51

7,9-Diethoxy DHP ((±)-7-ethoxy-9-ethoxy-6,7-dihydro-5H-pyrrolizine) 51

Toxicity of Sulfur-Conjugated Metabolites of PAs, in vivo 51

7-CYSDHP 51

7-GSDHP 52

Chemistry of Sulfur-Bound Pyrroles 52

Preparation and activation ofThiopropyl-Sepharose 6B resin 52 Recirculating aqueous flow system 53

Identification ofThiopropyl-Sepharose resin-bound pyrroles 54 Identification of liver sulfur-bound pyrroles 54 Quantitation of DHA binding to Thiopropyl-Sepharose resin 55

Studies of DHA Reactivity 56

Alkylation ofDHAs by 4-(p-nitrobenzyl)pyridine 56 Potentiometric measurement of hydrolysis ofDHAs 56

Inhibition of Hexokinase, in vitro 57

Yeast hexokinase assay 57 Rat brain hexokinase assay 58

Page 11: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

7

TABLE OF CONTENTS - Cont.

Gliicose-6-phosphate dehydrogenase assay 59

RESULTS 62

High-Speed Counter-Current Chromatography 62

Toxicity of Sulfur-Conjugated Pyrroles, in vivo 69

7-CYSDHP 70 7-GSDHP 70

Chemistry of Suifur-Bound Pyrroles 77

Identification ofThiopropyl-Sepaharose resin-bound pyrrole 77 Chemistry of liver sulfur-bound pyrroles 78 Quantitation ofDHA binding to Thiopropyl-Sepharose resin 78

Hydrolysis of DHAs 80

Alkylation of 4-(p-nitrobenzyl)pyridine by DHAs 83

Inhibition of Hexokinase by DHAs, in vitro 87

Yeast hexokinase 87 Rat brain hexokinase 88

Glucose-6-phosphate dehydrogenase 89

DISCUSSION 98

Counter-Current Chromatography 99

Toxicity of 7-CYSDHP and 7-GSDHP 104

Chemistry of Sulfur-Bound Pyrroles 107

Identification of Sulfur-Bound Pyrrole Ill

Relative Reactivity of DHAs Toward Thiopropyl-Sepharose 112

Hydrolysis of DHAs 113

Alkylation of 4-(p-nitrobenzyl)pyridine by DHAs 115

Inhibition of Yeast and Rat Brain Hexokinase by DHAs, in vitro 117

The Relationship Between the Physicochemical Properties of

DHAs and Their Further Metabolism and Toxicity 121

Macrocyclic diester PAs 121

Open ester DHAs 124

Conclusion 127

Page 12: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

8

TABLE OF CONTENTS - Cont.

APPENDIX A 129

Electron Impact-Mass Spectra of PAs 129

REFERENCES 145

Page 13: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

9

LIST OF FIGURES

! Structure and numbering system of macrocyciic diester PAs 21

2 Structure of open mono- and diester PAs 22

3 Structure of PA metabolites and derivatives 23

4 Pathways of PA metabolism 30

5 Mechanism of metabolic dehydrogenation of PAs to DHAs by

cytochrome P450 31

6 Proposed mechanism of alkylation by DHAs 32

7 Schematic of recirculating aqueous flow system for "trapping" of DHAs 61

8 Separation of PAs from Amsinckia tessellata using high-speed

counter-current chromatography 65

9 Separation of PAs from Symphytum spp. using high-speed

counter-current chromatography 66

10 Separation of alkaloids from Senecio douglasii var. longilobiis using

high-speed counter-current chromatography 67

11 Separation of PAs from Trichodesma incanum using high-speed

counter-current chromatogaphy 68

12 Representative potentiometric recording of the hydrolysis of

dehydroretrorsine and dehydrotrichodesmine 81

13 Rates of alkyaltion of macrocyciic DHAs by 4-(p-nitrobenzyl)pyridine 85

14 Rates of alkylation of open-ester DHAs by 4-(p-nitrobenzyl)pyridine 86

15 Concentration-effect curve of the inhibition of yeast hexokinase by dehydrotrichodesmine 91

16 Concentration-effect curve for the inhibition of yeast hexokinase by dehydrotrichodesmine and dehydrolycopsamine 92

17 Saturation curves for the inhibition of yeast hexokinase by dehydrotrichodesmine... .93

18 [Substrate/y versus [Substrate] plot of the inhibition of yeast hexokinase by

dehydrotrichodesmine 94

19 Effect of ATP and mannose on the inhibition of yeast hexokinase by dehydrotrichodesmine 95

20 Effect of DHAs on rat brain hexokinase, in vitro 96

21 Effect of dehydrotrichodesmine and dehydrolycopsamine on rat brain hexokinase,

in vitro 97

22 Proposed reaction of DHAs and Thiopropyl-Sepharose resin 109

23 A) Proposed mechanism of cleavage of sulfur-bound pyrroles by

mercuric chloride or silver nitrate. B) Structures of pyrrolic ethyl ethers

formed by solvolysis of sulfur-bound pyrroles by ethanolic silver nitrate 110

24 Proposed reaction between DHAs and 4-(/;-nitrobenzyl)pyridine 115

Page 14: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

10

LIST OF FIGURES - Cont.

25 Structure of 2,3-dihydro-l^-pyrroIizine-l-oI and dehydrosupinidine 116

26 Schematic of the hexokinase assay 118

27 Electron impact-mass spectrum of monocrotaline 128

28 Electron impact-mass spectrum of trichodesmine 130

29 Electron impact-mass spectrum of incanine 131

30 Electron impact-mass spectrum of retrorsine 132

31 Electron impact-mass spectrum of seneciphylline 133

32 Electron impact-mass spectrum of senecionine 134

33 Electron impact-mass spectrum of florosenine 135

34 Electron impact-mass spectrum of lycopsamine/intermedine 136

35 Electron impact-mass spectrum of 3-acetyl-lycopsamine/-intermedine 137

36 Electron impact-mass spectrum of 7-acetyl-lycopsamine/-intermedine 138

37 Electron impact-mass spectrum of 3,7-diacetyl-lycopsamine/-intermedine 139

38 Electron impact-mass spectrum of symphytine and isomers 140

39 Electron impact-mass spectrum of 7-ethoxy DHP 141

40 Electron impact-mass spectrum of 9-ethoxy DHP 142

41 Electron impact-mass spectrum of 7,9-diethoxy DHP 143

42 Positive ion FAB mass spectrum of 7-CYSDHP 144

Page 15: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

11

LIST OF TABLES

1 TLC Conditions for PAs and Pyrroiic Derivatives 60

2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using

High-Speed Counter-Current Chromatography 64

3 Summary of Wet Organ Weight / Body Weight Ratios in Rats at 21 Days

Following a Single Injection (i.v.) of 7-CYSDHP 71

4 Summary of Dry Organ Weight / Body Weight Ratios in Rats at 21 Days

Following a Single Injection (i.v.) of 7-CYSDHP 72

5 Summary of Wet Organ Weight / Dry Organ Weight Ratios in Rats at 21

Days Following a Single Injection (i.v.) of 7-CYSDHP 73

6 Summary of Wet Organ / Body Weight Ratios in Rats at 21 Days

Following a Single Injection (i.v.) of 7-GSDHP or Monocrotaline 74

7 Summary of Dry Organ / Body Weight Ratios at in Rats at 21 Days

Following a Single Injection (i.v.) of 7-GSDHP or Monocrotaline 75

8 Summary of Wet Organ / Dry Organ Weight Ratios in Rats at 21 Days

Following a Single Injection (i.v.) of 7-GSDHP or Monocrotaline 76

9 Relative Quantity of Pyrrole Bound to Thiopropyl-Sepharose Resin After

Exposure to DHAs 79

10 Rate Constants and Half-Lives of Hydrolysis of DHAs 82

11 Rate Constants for the Alkylation of DHAs by 4-(p-nitrobenzyl)pyridine 84

12 Effect of Dehydrotrichodesmine on Yeast Hexokinase Phosphotransferase

Activity in the Presence of Increasing Substrate Concentrations 90

13 Effect of Mannose (10 mM) and ATP (10 mM) Pretreatment on the Inhibition

of Yeast Hexokinase by Dehydrotrichodesmine 90

14 Relationship Between Pattern of Metabolism of PAs in the Isolated Perfused

Rat Liver and DHA Half-Life 125

15 Tissue-Bound Pyrroiic Metabolites 24 hr After i.p. Injection of PAs in Rats 126

Page 16: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

12

ABSTRACT

In humans, livestock and experimental animals, pyrrolizidine alkaloids (PAs) are

toxic as a consequence of their hepatic metabolism to reactive pyrrolic esters, or

dehydroalkaloids (DHAs). Despite their similarity in structures, PAs often vary markedly

in their lethality (LDjgS) and in the organs in which toxicity is expressed. We have

examined whether there are differences in the physicochemical properties of certain DHAs

which are associated with differences in patterns of metabolism and toxicity produced by

the parent PA. Using a potentiometric method to measure hydrolysis, it was determined

that the half-lives of the corresponding DHAs of retrorsine, seneciphylline, monocrotaline

and trichodesmine were 1.06, 1.60, 3.39 and 5.36 sec, respectively. These values were

supported by similar results from experiments measuring reactivity of DHAs toward 4-{p-

nitrobenzyOpyridine. Studies from the isolated rat liver perfused with PAs show that DHA

stability is related to patterns of metabolism and toxicity. Perfusion of the primarily

hepatotoxic retrorsine and seneciphylline is associated with a greater proportion of

metabolite released as non-toxic 7-glutathionyl-6,7-dihydro-l-hydroxymethyl-5/f-

pyrrolizine (7-GSDHP), a greater proportion alkylating liver macromolecules, and a lower

proportion released as DHA into the circulation. Perfusion with monocrotaline and

trichodesmine, PAs producing extrahepatic toxicity, produced lower proportions of 7-

GSDHP release and liver alkylation, and higher proportions of DHA released into the

circulation.

Other studies characterizing DHAs included the use of an in vitro enzyme assay in

which DHAs were shown to inhibit the phosphotransferase activity of yeast and rat brain

hexokinase. Parent PAs, and the hydrolysis product of DHAs, (±)-6,7dihydro-I-

hydroxymethyl-5f/-pyrrolizine (DHP) did not affect enzyme activity. In vivo smdies in

rats have established that glutathione and cysteine-conjugated pyrrolic metabolites of PAs

Page 17: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

13

likely represent detoxication pathways, providing further support for DHAs as the primary

toxic metabolite. We have examined the chemical form of suifur-bound pyrroles to

establish the importance of the 7-ester position in PA toxicity. Additionally, we have

developed an efficient technique for the rapid separation and purification of large quantities

of PAs using high-speed counter-current chromatography.

Page 18: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

14

INTRODUCTION

Pyrrolizidine alkaloids (PAs) are a large and structurally similar group of plant

compounds, most of which are toxic to animals, including man. PAs are fascinating tools

for scientific study for many reasons. PAs are important in plant chemosystematics (Kelley

and Seiber, 1992) and they are studied for their role in insect chemical defense and as

precursors to the biosynthesis of pheromones in moths and butterflies (Boppre, 1986). To

the toxicologist, PAs are of special interest because they require hepatic metabolism to

produce toxicity. Additionally, they are ideal for structure-activity studies because despite

similar structures, they sometimes produce markedly different toxicities.

PAs represent a significant world-wide public health risk. Human exposure results

from the use of PA-containing plants as food, herbal medicines or in beverages (Huxtable,

1989). Liver disease is the most common effect produced in humans, but extrahepatic

effects are observed in experimental animals and livestock. Human poisoning epidemics

have occurred in several countries due to contamination of cereal grains with PA-containing

plants. PAs are probably the most significant plant toxins in terms of their ill effects to the

health of man and animals, and their associated economic costs (Huxtable, 1989). The

continued study and dissemination of knowledge concerning PAs is an important step

towards increasing awareness and subsequent reduction of PA poisonings.

Toxicity of PAs is related to their metabolism to reactive pyrrole intermediates,

which can bind to cellular constituents (Mattocks, 1986; Glowaz et al., 1992). The

relationship between PA structure and toxicity is both complicated and poorly understood.

Some PAs such as retrorsine are generally hepatotoxic, while monocrotaline is

pneumotoxic as well. Trichodesmine is a potent neurotoxin. How structurally similar

alkaloids produce such different toxicities is unclear. We have addressed this question by

investigating the chemical properties of a series of PA metabolites, and how their respective

Page 19: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

15

structures influence their stability, patterns of metabolism and toxicity. The results provide

further information on the relationship of hepatic metabolism of xenobiotics and tlie

extrahepatic toxicity they produce.

Plant Sources of PAs

More than 350 PAs have been identified, from over 300 plant species (Huxtable,

1989; Logie et al., 1994). This represents approximately 3% of the worlds flowering plant

species. PA containing genera have worldwide distribution. Some important sources are

the genus Crotalaria (Leguminosae), the tribe Senecionae (Asteraceae), and the

Boraginaceous genera. Amsinckia, Heliotropium, Symphytum and Trichodesmine (Smith

and Culvenor, 1981; Mattocks, 1986). Many plants contain both PAs and a variable

proportion of the corresponding A'-oxide. Quantitative and qualitative PA content of a

particular plant may be influenced by its location, time of year, or other environmental

factors.

PA Exposure

Humans are exposed to PAs because of either intentional or unintentional

consumption of plant products containing them. These products may be consumed in the

form of herbal medications or teas, as green vegetables, or as grain contaminated with PA

containing seeds. PAs have been found in honey produced by bees feeding on the nectar of

plants (Deinzer et al., 1977; Culvenor et al., 1981) containing them, and in the milk of

goats and dairy cows fed on ragwort {Senecio jacobaea) (Dickinson et al., 1976;

Dickinson, 1980). However, there are no known cases of human toxicity resulting from

the consumption of contaminated milk or honey (WHO, 1988). Chronic and heavy herb

users, young children, elderly, and sick individuals are particularly susceptible to

intoxication (Huxtable, 1990b). Additionally, certain ethnic groups such as Mexican-

Page 20: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

16

Americans and Native Americans are at risk to PA exposure because of extensive traditional

use of herbal remedies. Prevention of poisoning lo humans can be achieved by controlling

PA containing plants in agricultural areas, by educating populations at risk, and by banning

the sale of PA containing herbal products.

Exposure to PAs by the use of herbal products in the form of infusions or medicinal

remedies occurs for several reasons. Ingestion may be intentional, as in the case of comfrey

(Symphytum) preparations, or accidental, such as when an herbal product is incorrectly

identified. Medicinal use of herbs in traditional medicine has long been practiced in many

countries including China, Greece, India and the United States (WHO, 1988). Many cases

of suspected PA poisonings have occurred in the United States and other countries as result

of such use (Huxtable, 1989). The incidences of these poisonings likely will continue in

light of the growing interest in herbal products as food and home remedies in the

developing world (WHO, 1988; Sperl etal., 1995).

PA-induced veno-occlusive disease of the liver was endemic in Jamaica and other

West Indian countries for many years (Hill and Rhodes, 1953; Bras et al., 1954; Stuart and

Bras, 1956; Stuart and Bras, 1957). This illness was associated with the consumption of

herbal or "bush" teas which often contained Crotalaria, in particular C.fidva, although

Senecio has been implicated as well (Bras et al., 1957). Herbal teas were frequently used

as infant remedies, and many children died of massive ascites associated with veno-

occlusive disease of the liver. The morphological changes to the liver brought about by this

disease were first recognized and described by Bras et al. (1954).

Comfrey (Symphytum spp.) is cultivated in many areas of the world for use as an

herbal tea, digestive-aid and cosmetic. These preparations are known to contain PAs and

are implicated in cases of PA poisoning (Ridker er a/., 1985; Huxtable, 1989;. PA levels in

comfrey products vary considerably, depending on the species and strain of Symphytum

and the plant part utilized. Comfrey root, which is frequently used for herbal teas, contains

Page 21: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

17

the highest concentrations of alkaloid. Long-term use of comfrey products is common

(Huxtabie eiai., 1986); thus effects of toxicity may appear after chronic use. In addition to

being hepatotoxic, some of the alkaloids from Symphytum are known to be carcinogenic in

rats (Hirono et ai, 1978, 1979). Comfrey products are still widely available in the United

States.

Humans may accidentally be poisoned by contamination of food grains with PA-

containing plant material, particularly seeds. In 1950, an outbreak of Trichodesma

poisoning occurred in the Samarkand region of Uzebekistan, killing at least 44 people

(Ismailov et al., 1970). More than 200 individuals were affected, due to the consumption

of bread prepared from grains contaminated with the seeds of T. incanum. Deaths were

attributed to central and peripheral neuropathies. Until recently, PA poisoning in India,

Afghanistan and the former USSR was endemic (Huxtabie, 1989) due to contamination of

grains with Heliotropium species. These outbreaks of PA poisoning are among the largest

known, with some 35.000 people exposed in one epidemic of liver disease in Afghanistan

(Mohabbat erfl/., 1976; Tandon era/., 1978a, 1978b).

PA poisoning of livestock is a serious economic problem in some areas of the

world (Mattocks, 1986). Many PA containing plants, such as Senecio jacobaea (ragwort)

and various Crotalaria thrive during drought or in pastures of poor condition. Thus they

may be eaten by stock during times of hardship. At one time, S. jacobaea was said to cause

more cattle losses than all other poisonous plants combined (Mattocks, 1986). Cattle

feeding on Senecio typically display liver toxicity characterized by hepatic lesions and

veno-occlusive disease (Thorpe and Ford, 1968; Johnson et al., 1985; Mattocks, 1986).

Livestock which have been grazing on Crotalaria frequently display extrahepatic

symptoms, especially to the lungs and kidneys (Mattocks, 1986).

Page 22: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

18

PA Isolation, Separation and Puriflcation

Extraction from plant material

PAs are extracted from either fresh or dried plant material in the form of ground

leaves, stems, seeds, etc. Plant material is generally extracted with hot alcohol in a Soxhlet

apparatus. The alcohol is evaporated and the residue is resuspended in dilute acid. This

solution is extracted with an organic solvent to remove chlorophylls and lipids. Zinc dust is

added to the acidic aqueous portion to reduce A/^oxides to their respective tert-hases. The

solution is made basic with ammonia and extracted with chloroform, yielding the alkaloidal

fraction. Extracts containing polar alkaloids such as lycopsamine are saturated with NaCl to

increase alkaloid extraction (Culvenor and Smith, 1966). If the fraction contains only one

PA, it may often be obtained pure by recrystallization from alcohol. Other alkaloidal

fractions are obtained crystalline, but are comprised of several PAs which must be

separated chromatographically (Adams and Govidachari, 1949). Fractions homAmsinckia

or Symphytum are typically isolated as brown gums (Culvenor and Smith, 1966; Roitman,

1983; Cooper et al., 1996), resisting all efforts at crystallization. Most fractions are isolated

as PA mixtures, often in the form of stereoisomers.

Spectrophotometric determination of PAs

Unsaturated PAs may be quantified spectrophotometrically by reacting with Ehrlich

Reagent after the PA has been converted to a pyrrole (Mattocks, 1967, 1968a). Initially, the

PA is oxidized with hydrogen peroxide to its respective A^-oxide which is then converted to

a pyrrole by acetic anhydride. Pyrroles react strongly with modified Ehrlich Reagent (4-

dimethylaminobenzaldehyde in acidified ethanol) (Figure 3) to produce a magenta solution

with a at 562 nm. This method of detection is sensitive (below 5 |ig for some PAs)

and reproducible (Mattocks, 1968a). PAs may be visualized on TLC plates by first

Page 23: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

19

spraying with methanolic o-chloranil, followed by Ehrlich Reagent. Pyrroles may be

visualized by Ehrlich Reagent alone. Both yield magenta colored spots when treated in this

manner.

Preparative chromatographic separation ofPAs

As only monocrotaline has been available commercially, it is often necessary to

isolate and purify large amounts of PAs for toxicological or other studies. Mixmres of PAs

have been separated and purified with varying degrees of success using column

chromatography with different types of adsorptive media. Typical matrices have been

alumina and silica (Segall, 1984; Mattocks, 1986). However, column chromatography is

slow, and repeated runs are often necessary for complete or partial resolution of alkaloids

(Adams and Govindachari, 1949). Preparative HPLC has recently been used for separation

of PAs (Segall, 1984). Columns employed are reversed phase C-18 or PRP types.

Separations are rapid, with good resolution, but columns are expensive and cannot separate

large quantities. PAs have been separated with droplet counter-current chromatography

(DCCC) (Zalkow era/., 1988; Asibal era/., 1989a, 1989b; Culvenorera/., 1980) with

excellent results, but the technique is slow and has generally been replaced by the newer

high-speed method described below.

High-speed counter-current chromatography

Counter-current chromatography (CCC) is a method of liquid-liquid partition

chromatography that does not employ a solid matrix. Ito et al. (1966) observed that two

immiscible liquids will become uniformly segmented in the coils of a helical column when

the latter is rotated. The liquid designated as the stationary phase is retained by centrifugal

force, while the mobile phase is pumped through die rotating column. Advantages of high­

Page 24: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

20

speed CCC include total sample recovery, wide choice of solvent systems, rapid separation

Limes, good resolution (350-1000 plates) and high reproducibility. Since there is no solid

support mechanism within die column, sample contamination, adsorption, and denaturation

are normally eliminated (Conway, 1990). A recently developed instrument for high-speed

CCC, the multilayer coil planet centrifuge (Ito et al., 1982), is well suited for the rapid

separadon and purification of natural products on a preparative scale (Conway, 1990;

Fischer a/., 1991; Zhang era/., 1988; Marston and Hostettman, 1994).

PA Structure

PAs are generally derivadves of 1-methylpyrrolizidine (Mattocks, 1986), consisting

of two fused five-membered rings sharing a common nitrogen at the 4 position (Fig. 3).

When hydroxylated, this structure is referred to as a necine. Necine bases may also be M-

methylated, forming otonecine. Most PAs have an acid moiety, termed a necic acid,

esterifying either or both the 7 or 9-hydroxyl position of necine. Hepatotoxic PAs are esters

of unsaturated necines, containing a double bond in the 1:2 position of the pyrrolizidine

ring. PAs may exist as monoesters, diesters, or as macrocyclic diesters. Figures 1 and 2

shows the structures and numbering system of PAs used in this dissertation. Many necic

acids can comprise the ester portion of PAs. These often differ slightly, and many exist as

stereoisomers. Differences in the acid structure can cause marked differences in toxicity

produced by the parent PA. The macrocyclic diesters such as retrorsine are generally the

most toxic. Open diester PAs such as symphytine or lasiocarpine may be as hepatotoxic as

some macrocyclics, but others are much less so (Mattocks, 1986). Monoester PAs such as

lycopsamine and indicine are weakly hepatotoxic in comparison to diester PAs (Fowler and

Schoental, 1967; Schoental, 1968).

Page 25: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

21

CH

3 5

R=H; Senecionine R=OH; Retrorsine

CH

Seneciphylline

18

R'=Me, R"=OH; Monocrotaline

R'=CH(CH3)2, R-=H; Incanine

R'=CH(CH3)2, R"= OH; Trichodesmine

O II

.CH

CH

Florosenine

Figure 1. Structure and numbering system of macrocyclic diester PAs. All alkaloids are retronecine based, except for florosenine, which contains an otonecine base. Hepatotoxic PAs are esters of unsaturated necines, containing a 1-2 double bond.

Page 26: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

22

I JZHs

O CH

^0-C-C-«OH H J I

\ R3.-C-CH:

V 'r-

R'=Me, R^=H, R^=H, R'^=OH; Symphytine

R'=H, R"=Me, R^=H, R'^=OH; Symlandine

R'=Me, R~=H, R^=OH, R'^=H; Myoscorpine

R'=:H, R~=Me, R^=OH, R'^=H; Echiumine

R'=H, R"=H, R^=0H; Lycopsamine

R'=H, R^=0H, R^=H; Intermedine

R'=H, R"=H, R^=AC; 3-Acetyl-lycopsamine

R'=H, R"=AC, R^=H; 3-AcetyI-intermedine

R'=AC, R-=H, R^=0H; 7-Acetyl-lycopsamine

R'=AC, R^=0H, R^=H; 7-Acetyl-intemiedine

R'=AC, R^=H, R^=AC; 3,7-Diacetyl-lycopsamine

R'=AC, R^=AC, R^=H; 3,7-Diacetyl-intermedine

HjC CH3 \ /

O CH

Figure 2. Structure of open mono- and diester PAs.

Page 27: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

CH

I -methylpyrrolizidine

R' = H, R~ = OH; Dehydroretronecine

R' = OH, R" = H; Dehydroheliotridine

COtH O O I ' II II I CH^NHC- CH- NHC-CH2CH2CHNH0

I

OH

7-GSDHP

OH HO

DHP

ROCO OCOR'

NMe CH

Ehrlich Adduct, 562 nm

O II

H-,NCHCOH

CH

OH

7-CYSDHP

Figure 3. Structure of PA metabolites and derivatives. PAs are derivatives of 1-methylpyrrolizidine. DHP is a racemic mixture of dehydroretronecine and dehydroheliotridine.

Page 28: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

24

Chemical Properties of DHAs

DHAs are formed in the iiver by the metabolic dehydrogenation of unsaturated PAs,

catalyzed by cytochrome P-450. The DHAs are also easily prepared chemically from PAs

by oxidation with the quinones o-bromanil or o-chloranil (Mattocks etal., 1990). The

subsequent pyrrolic product differs from the PA in that it contains an additional double

bond in the necine ring (Figure 4). Because the DHA is much more reactive than the parent

PA, they are characterized by their susceptibility to hydrolysis and polymerization under

acidic conditions and rapid reactivity toward nucleophiles (Culvenor etal., 1970). This

makes DHAs difficult to work with since special handling is required. Their inherent

reactivity is also why they have never been isolated intact from animals exposed to PAs or

from in vitro systems such as liver microsomes. DHAs are reactive because the conjugated

double bonds within the pyrrole ring leads to a dispersion of charge, made possible by the

resonance donation of pi electrons from the nitrogen atom. This conjugation makes the

esters linked to the pyrrole at C7 and C9 labile, hence they become good leaving groups. In

an aqueous environment, this instability leads to rapid hydrolysis and formation of a more

stable pyrrolic alcohol, (±)-6,7-dihydro-7-hydroxy-I-hydroxymethyl-5//-pyrrolizine

(DHP) (Kedzierski and Buhler, 1986). The extent of racemization at C7 when DHAs react

with alcohols or water indicates hydrolysis proceeds through an S^l mechanism, via

formation of a pyrrole-stabilized carbonium ion (Figure 5) (Mattocks, 1986; Kedzierski and

Buhler, 1986). Thus DHP is a racemic mixture of dehydroheliotridine and its enantiomer,

dehydroretronecine. Diester PAs are potentially bifunctional alkylating agents because they

contain a ester at both C7 and C9. The C7 carbonium ion is more reactive than C9 as it

forms a more stable secondary carbonium ion as opposed to a primary carbonium ion

formed at C9. Thus, once substitution has occurred at C7, a similar reaction can occur at

C9. Formation of stabilized carbonium ions allows covalent reactions with cellular

nucleophiles such as glutathione or macromolecules (Figure 5). The macrocyclic DHAs

Page 29: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

25

have been shown to cross-link DNA (Petry et al., 1984; Hincks et al., 1991), presumably

due to their bifunctionality. DHP is an amino alcohol and is considerably more stable than

DHAs. This is because hydroxyl functions do not make good leaving groups, despite the

resonance stabilization of the pyrrole ring. Under acidic conditions however, DHP can

react with nucleophiles (Mattocks and Bird, 1983b) and can alkylate DNA (Petry etai,

1984; Hincks era/., 1991).

PA Metabolism

PA must be metabolized to produce toxicity. Only the liver appears capable of

bioactivation of PAs to toxic metabolites. Several lines of evidence, reviewed by Mattocks

(1986), support the idea that PA metabolites, and not the parent PAs, are responsible for

toxicity: (1) PAs are not toxic at injection sites, nor are they toxic to the skin; (2) the liver

is the main target organ of toxicity, regardless of route of administration; (3) drug

treatments which modify metabolism influence the toxicity of PAs; 4) animals vary greatly

in susceptibility to PAs; 5) PAs are rather inert compounds, and they would not be

expected to react with cellular constituents under physiological conditions; 6) many insects

accumulate large amounts of PAs in their tissues without harm.

In the laboratory animal, the main routes of metabolism for PAs are ester

hydrolysis, yV-oxidation and dehydrogenation to pyrrolic products (Figure 4). The first two

pathways represent detoxication mechanisms, while the latter is associated with toxicity.

Additionally, much alkaloid is excreted unchanged, principally in the urine (Eastman et al.,

1982). The balance of these pathways will determine the eventual biological effect

produced by a PA in a particular animal (Mattocks, 1986). The toxic properties of pyrrolic

metabolites are consistent with the cytotoxic action of PAs, and it is improbable that other

important routes of bioactivation will be discovered (Mattocks, 1986).

Page 30: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

26

N-Oxidation

Of the two major oxidation products formed by the hepatic metabolism of PAs,

A^-oxidation represent a detoxication pathway. Many species metabolize PAs to their

corresponding N-oxides, including rats, mice, hamsters, guinea pig, sheep, quail, fowl,

and humans (Mattocks, 1968; White et al., 1973; Williams et al., 1989). PAs may be

converted to ^/-oxides by both P450 and flavin-containing monooxygenase (FMO), with

the requirement for and NADPH. The relative contribution by either enzyme is species

and tissue dependent. Purified pig liver FMO has been shown to oxidize senecionine to its

A/^oxide, while in rat liver, P450 makes the most significant contribution (Williams et ai.

1989). In humans, P4503A4 is the major catalyst for the conversion of senecionine to its

A^-oxide (Miranda et al., 1991). Guinea pig liver, lung, and kidney microsomes convert

senecionine primarily to A'-oxide via FMO, with little pyrrole activation observed (Miranda

et al., 1991). This is in keeping with the resistance to PA intoxication by guinea pigs. The

A/-oxides are water soluble and are rapidly excreted in the urine (Mattocks, 1968a).

Additionally, they do not appear to be a substrate for pyrrole formation (Jago et al., 1970:

Mattocks and White, 1971). Although A/-oxides may be reduced to pyrrolizidine bases in

the liver, this appears to be a minor pathway (Powis and Wincentsen, 1980). A/^oxides are

extensively reduced in the gut however, and the alkaloid base is reabsorbed. Rats, sheep

and rabbits all have the capacity to reduce A^-oxide PAs (Mattocks, 1971; Powis et ai,

1979). The reduction of A^-oxides to tertiary bases in the gut is made evident by their higher

toxicity when administered orally rather than i.p. (Mattocks, 1972).

Hydrolysis

PAs are subject to enzymatic hydrolysis by hepatic esterases (Mattocks, 1972,

1982; Dueker et al., 1992, 1995; Chung and Buhler, 1995). This results in the formation

Page 31: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

27

of a necine base and necic acid, neither of which are known to be toxic (Culvenor et al.,

1976). Hydrolysis is tlius considered a detoxication pathway for PAs. Hydrolysis of PAs

is markedly reduced in rats pretreated with tri-orthocresylphosphate, an esterase inhibitor,

and subsequently formation of toxic pyrrolic metabolites is increased (Mattocks, 1981;

Mattocks, 1986). Both chemical (base hydrolysis) and enzymatic hydrolysis (rat liver

homogenates) of natural PAs and synthetic derivatives are strongly influenced by steric

hindrance from the acid substitution in the vicinity of the ester groups (Mattocks, 1982).

PAs containing esters with short unbranched chains such as acetate and propionate are

subject to rapid hydrolysis both chemically and enzymatically. Open branched diesters,

such as retronecine ditiglate, or macrocyclics like monocrotaline and retrorsine are much

more resistant to hydrolysis. Hepatic esterase hydrolysis of PAs may be a relatively minor

metabolic pathway in rats, but appears be significant in guinea pigs (Chung and Buhler,

1995). The contribution of esterases in blood and other tissues toward PA hydrolysis is

unknown. Guinea pigs are especially resistant to PA toxicity (White et al., 1973; Mattocks,

1986; Miranda et al., 1991). However, they are particularly sensitive to the PA jacobine

(Swick et al., 1982). Recent studies have shown that purified guinea pig carboxylesterases

GPHl and GPLl are weakly active toward jacobine in comparison to other PAs (Chung

and Buhler, 1995; Dueker et al., 1995). In the whole animal, more jacobine may be

available for bioactivation to DHA, since hydrolysis of the alkaloid appears limited.

Dehydrogenation to pyrroles

PAs are metabolized in the liver to pyrroles by the cytochrome P450s, an enzyme

system responsible for the detoxication of many foreign compounds. These heme-

containing enzymes catalyze oxidative metabolism by the transfer of an oxygen atom or

oxidative equivalent into a variety of xenobiotics. Occasionally, this metabolism results in

Page 32: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

28

the generation of reactive, electrophilic compounds, a process termed "bioactivation"

(Vamvakas and Anders, 1990). Tne toxicity of many drugs, chemicals and pollutants

results from the formation of reactive intermediates during their metabolism.

The P450s exist as isozymes in a given animal species. In rats and man, the

P4503A isoform is responsible for the majority of pyrrole formation (Williams et al., 1989:

Miranda et al., 1991), while P4502B appears to be the most significant isoform in the

guinea pig (Chung et al., 1995). Pyrrole formation is associated with the cytotoxicity of

PAs (Mattocks, 1968; Mattocks and White, 1971; White et al., 1973). As proposed by

Mattocks, (1968) PAs are initially dehydrogenated to pyrrolic esters, or DHAs (Figure 6).

Dehydrogenation is believed to result from hydroxylation at C8 on the unsaturated necine

moiety (Mattocks and White, 1971; Mattocks and Bird, 1983), forming a chemically

unstable carbinolamine. This species would spontaneously decompose, losing first OH"

and then H"", to yield the DHA (Figure 6) (Mattocks, 1986). Once a DHA has formed, it

can alkylate macromolecules and soluble nucleophiles such as glutathione or rapidly

hydrolyze to DHP. This compound was identified as the major urinary metabolite in rats

given i.p. monocrotaline (Hsu etal., 1973), providing indirect evidence for the formation

of dehydromonocrotaline. DHP is also produced from PAs using in vitro metabolism

systems (Mattocks, 1968; Jago etal., 1970; Glowaz etal., 1992). Relative pyrrole levels in

the liver following exposure to different PAs correlate well with the acute hepatotoxicity of

the parent alkaloids (Mattocks, 1972). The toxicity of the DHAs, and their hydrolysis

product, DHP, is discussed below.

Toxic Effects of PAs

PA toxicity is a consequence of their hepatic metabolism to pyrroles. These

metabolites are electrophilic and bind covalently to tissue nucleophiles, either in the liver or

Page 33: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

29

at more distant sites sucli as the lungs. The covalent modification of critical cellular

macromolecuies, including proteins and nucieic acids, presumably iniuates ceil damage.

Hepatotoxicity

The primary acute action of PAs in laboratory animals and humans is

hepatotoxicity, producing a pathognomonic syndrome known as veno-occlusive disease of

the liver (Bras et ai, 1954). Acute poisoning in animals leads to zonal hemorrhagic

necrosis followed by a narrowing of the endothelium of the small hepatic veins leading to

occlusion. Hepatic sinusoids may become dilated and pooled with blood, a finding

resembling the human condition. HepatocytomegaJy and fibrotic lesions are characteristic

pathologies observed in animals with chronic PA exposure. Megalocytes are produced as a

result of the antimitotic activity of PA metabolites on parenchymal cells. Interestingly,

megalocytes have not been observed from the liver of humans exposed to PAs (Mattocks et

ai, 1986). Hepatic veno-occlusive disease in man due to PA consumption has been well

documented (Hill and Rhodes, 1953; Bras et al., 1954; Bras and Hill, 1956; Stuart and

Bras, 1957). The acute human condition is characterized by portal hypertension with

massive ascites attributable to occlusion of centrilobular or sublobular hepatic veins (Sped

et al., 1995). The patient (most often a child) experiences rapid onset of abdominal

discomfort, with a smooth and firm hepatomegaly (Stuart and Bras, 1957). Abdominal

swelling due to ascites and dilated veins in the abdominal wall are observed. Death may

shortly ensue, but veno-occlusive disease is often reversible if PA exposure is stopped and

the patient given conservative therapy (Bras and Hill, 1956, Lyford et al., 1976; Sperl et

al., 1995). The chronic condition is characterized by cirrhosis, which is clinically

indistinguishable from other forms of cirrhosis. Jaundice may be observed and there is

biochemical evidence of liver dysfunction.

Page 34: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

30

ROCO ROCO HO OH

Hydrolysis

P-450, FMO Esterases

PA

PA A'-oxide

Dehydrogenation P-450

Retronecine

HO ROCO OH GS OH

Hydrolysis GSH •Nv

DHP

-N^

DHA

I GSH \

7-GSDHP

Alkylation of cellular nucleophiles

HEPATIC AND EXTRAHEPATIC TOXICITY

Figure 4. Pathway^f PA metabolism. Cytochrome P-450s, flavin-containine esterases catalyze the metabolism of PAs Toxicity is a result

of dehydrogenation by cytochrome P-450 to a reactive pyrrole intermediate, th^DF^

Page 35: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

31

OCOR"

P-450

Liver

OCOR

-OH

OCOR" RO OCOR" RO

H

Figure 5. Mechanism of metabolic dehydrogenation of PAs to DHAs by hepatic cytochrome P-450, as proposed by Mattocks and Bird (1983). The first step in bioactivation is the hydroxylation of the pyrrolizidine ring at C8. This results in an unstable carbinolamine which decomposes to a dihydropyrrolizine (DHA) by losing first OH" followed by H"^.

Page 36: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

32

CHiO- C- R-R ' — C - 0 Nu CH^O-C- R-

Nu

O. 11^ , CH2O- C- R- Nu Nu Nu

Nu

Nu = nucleophile

Figure 6. Proposed mechanism of alkylation by DHAs. Esters linked to the pyrrole ring are made labile by their proximity to nitrogen. This structure favors the formation of pyrrole stabilized carbonium ions at C7 and C9, which proceeds through an Sjsjl mechanism. The positively charged pyrrole nucleus can then react with cellular nucleophiles, likely producing the toxicity associated with PAs. From Huxtable (1990).

Page 37: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

33

Pneiimotoxicity

Many PAs effect the lungs of certain animals, producing pulmonary arterial

hypertension and right ventricular hypertrophy. This subject has been reviewed by

Huxtable (1990). The 11-membered macrocyclic diesters such as fulvine and

monocrotaline are the most active in producing lung damage, although the Senecio alkaloid

seneciphylline was reported to be pneumotoxic in one instance (Ohtsubo et al., 1977).

Monocrotaline is often used to produced the syndrome in animal models due to the

similarity to the human condition (Kay and Heath, 1969; LaFranconi and Huxtable, 1981).

In the rat, the appearance of pulmonary hypertension following exposure to a single dose

of monocrotaline is delayed for up to 2 weeks despite the appearance of liver damage. The

pneumotoxic process is initiated upon exposure to the alkaloid, which is largely excreted

and metabolized by 24 hours (Candrian et al., 1985; Huxtable, 1990). Following exposure

to the pneumotoxic PA, a remodeling of the pulmonary vasculature occurs associated with

endothelial hyperplasia, arterial medial thickening, spreading of arterial smooth muscle into

unmuscularized areas, smooth muscle hypertrophy and increased lung mass. Platelet

accumulation initiates the formation of microdirombi in the capillaries, which leads to

capillary occlusion and edema. Row resistance and thus arterial pressure is increased,

forcing a greater workload upon the right heart. This in turn produces right heart

hypertrophy and cor pulmonale. The degree of pulmonary arterial hypertension produced

by monocrotaline strongly correlates with the degree of right heart hypertrophy (Kay and

Heath, 1969; Huxtable etal., 1978; Huxtable, 1990). Many complex biochemical changes

occur during PA-induced pulmonary hypertension, including alterations in the metabolism

of the vasoactive substances serotonin and angiotensin (Huxtable etal., 1978; Huxtable.

1979; LaFranconi and Huxtable, 1983). Increases in the synthesis of protein and RNA

occur both in the right heart and the lung (LaFranconi et al., 1984).

Page 38: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

34

Neurotoxicity

Neurotoxicity is observed in some instances of PA poisoning of livestock,

especially in horses (Rose et ai, 1957a, 1957b). The non-specific effects reported were

attributed to a hyper-ammonemia following liver disease. Spongy degenerations in the

central nervous systems in cattle, sheep and pigs have been reported in animals exposed to

PAs (Hooper ef a/., 1974; Hooper, 1975a, 1975b).

The Trichodesma alkaloids produce neurotoxicity as their primary action (WHO,

1988; Huxtable, 1989), as opposed to the hepatotoxicity associated with other PAs. The

neurotoxicity of Trichodesma in dogs, rabbits and mice has been reviewed in the Russian

literature (Shtenberg et al., 1955; Ismailov et al., 1970). Mice treated with Trichodesma

alkaloids develop hind limb paralysis, opisthonus and clonic convulsions. An equine

disease known as "suiljuk" is associated with consumption of T. incanum by horses

(Smimov etal., 1959a, 1959b, Smimov and Stoljarova, 1959). Degenerative changes in

liver, lung, kidneys and heart were reported. Cows are similarly effected (Smimov et al.,

1959b). An outbreak of human PA poisoning due to T. incanum contaminated grain

occurred in the Samarkand region of Uzebekistan from 1942-1951 (Ismailov etal, 1970).

The disease, known as "Ozhalanger encephalitis", affected over 200 adults. Curiously,

children under the age 10 were not affected. Observed symptoms, appearing after a 10 day

latency, included nausea and vomiting, malaise, headaches, delirium, loss of

consciousness, pathological reflexes and paresis of the extremities and facial nerves. Forty-

four of the affected individuals died. T. incanum contains the Crotalaria type PAs incanine

and trichodesmine (Yunosov and Plekhanova, 1957a, 1957b, 1959), which are structurally

similar to monocrotaline (Figure 1).

Page 39: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

35

Carcinogenicity and anti-tiimor activity

PA are among the few compounds found in plants that are carcinogenic. This is not

surprising in light of the powerful in vitro antimitotic and mutagenic effects of many PAs

on cells. While increased incidence of cancer or mmors in humans during PA poisoning

outbreaks have not been reported, PAs are moderately carcinogenic in experimental

animals (Mattocks, 1986). A variety of unsaturated PAs have been examined for

carcinogenicity in rats. Regardless of the dosing schedule (continuous or intermittent) and

route of administration, tumors occurred most commonly in the liver (Mattocks, 1986).

Thus it is likely that hepatic metabolism plays a key role in carcinogenicity. Both

macrocyclic PAs {Senecio type) (Harris and Chen, 1970) and open esters (Hirono et al.,

1978) are hepatocarcinogens. DHP is also carcinogenic, but less so than monocrotaline

(Mattocks, 1986), indicating that other PA metabolites, such as the DHAs, likely play a

significant role in carcinogenicity.

Some PAs and PA A^-oxides have been shown to posses antitumor activity.

Indicine A'^oxide has demonstrated antitumor activity in mice against Walker 256

carinosarcoma (Kugelman etal., 1976). This compound has undergone clinical trials in

humans against childhood leukemia (Miser et al., 1992) and advanced solid tumors

(Kovach etal., 1979). Remissions were achieved in several cases of the leukemias, but the

hepatotoxicity produced as a side effect precludes the continued pursuit of indicine AZ-oxide

as an anticancer drug.

Mutagenicity and cytotoxicity

PAs possessing the structural features necessary to cause hepatotoxicity also can

alkylate DNA. The pyrrolic metabolites dehydroheliotridine and dehydroretronecine can

alkylate DNA as well (Black and Jago, 1970; Robertson, 1982). Robertson (1982) has

Page 40: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

36

identified a 7-adduct of dehydroretronecine bound to the N2 position of deoxyguanosine.

Tlie DHAs, dehydromonocrotaline and dehydroretrorsine, were much more active cross-

linkers of DNA than dehydroheliotridine (White and Mattocks, 1972). This is in keeping

with the much greater electrophilic reactivity of the former.

In mutagenicity assays, metabolic activation of PAs is necessary to produce

maximum effect, as DNA damage is probably caused primarily by the DHA. Bifunctional

alkylating PAs (diesters) are more effective mutagens, and can cross-link DNA (Petry et

al., 1984; Hincks et al., 1991). Metabolically generated DHAs containing an a, p-

unsaturation, such as dehydrosenecionine, were more active at cross-linking DNA in

cultured bovine kidney epithelial cells than the cyclic diester dehydromonocrotaline or open

ester DHAs (Kim et al., 1993). This may be due to extent of metabolic activation of

different PAs rather than alkylating ability of the reactive species. PAs are mutagenic in

Drosophila melanogaster. Salmonella typhimuriiim and CHO cells (Ord et al., 1985;

Mattocks, 1986; Carballo er a/., 1992; Frei etai, 1992). Mutagenicity in Drosophila is well

correlated with hepatotoxicity of PAs in rodents. Frei et al. (1992) have shown that

macrocyclic diester PAs were the most genotoxic to Drosophila, followed by open diesters,

and open monoesters. The C9 monoester supinine, which lacks an alkylating function at

C7, was not genotoxic.

Most hepatotoxic PAs, and the metabolite DHP, produce a powerful antimitotic

effect on liver parenchymal resulting in characteristic enlarged hepatic parenchymal

(megalocytes) cells (Peterson, 1965; Jago, 1969; Mattocks and Legg, 1980). Mitosis is

persistently inhibited during the S or early G phase of the cell cycle, possibly by inhibiting

the synthesis of specific proteins required for cells to progress from G, to mitosis (Jago

and Samuel, 1975). Mattocks and Legg (1980) observed a similar effect in vitro using a rat

liver parenchymal cell line with a series of pyrrolic alcohols, including DHP. In the

presence of rat liver S9, PAs with an a, [B-unsamration {Senecio alkaloids) were the most

Page 41: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

37

effective in producing megalocytosis in bovine kidney epithelial cells (Kim etal., 1993).

Monocrotaline, and open diesters produced only a weak response.

Metabolites Responsible for Extrahepatic Toxicity

Extrahepatic toxicity is likely a result of hepatic metabolism, as other tissues,

including the lungs, appear incapable of PA bioactivation (Mattocks, 1986; Huxtable,

1990; Miranda et al., 1991). The reason that some PAs cause extensive extrahepatic

toxicity is still poorly understood. Whether extrahepatic toxicity is due to the highly reactive

DHA or a stabilized secondary pyrrolic metabolite is uncertain.

DHAs

DHAs are the primary pyrrolic metabolites produced in by liver from PAs. DHAs

are highly reactive, alkylating a variety of cell nucleophiles, including proteins and DNA

(Petry etal., 1984; Hincks etal., 1991; Kim etal., 1995), and are likely responsible for

hepatic toxicity. DHAs injected into mesenteric veins produce hepatotoxicity similar to the

parent alkaloids (Mattocks, 1986), but much lower doses of the former are required.

Dehydromonocrotaline, given via tail vein to rats in a non-protic solvent, produces

pneumotoxicity consistent with the parent alkaloid (Butler et al., 1970; Plestina and Stoner,

1972; Bruner et al., 1983). Doses as low as I mg/kg of the former are required compared

to approximately 60 mg/kg for monocrotaline (Pan et al., 1993). However, the role of

DHAs in extrahepatic toxicity has remained unclear. They are extremely reactive, having

aqueous half-lives of only a few seconds (Mattocks and Jukes, 1990; Cooper and

Huxtable, 1996). It has thus been much debated whether DHA produced in the liver could

survive transport to the lungs, or organs even further downstream (Huxtable, 1990; Pan et

al.. 1993).

Page 42: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

38

Sulfur conjugated pyrroles: 7-GSDHP and 7-CYSDHP

Phase n conjugation with the tripeptide glutathione (y-gltamyl-L-cysteinyl-glycine)

is generally known as a detoxication pathway, although several classes of xenobiotics form

reactive intermediates by this route (Anders etal., 1988; Vamvakas and Anders, 1990;

Koob and Dekant, 1991). Large quantities of sulfiar-conjugated pyrroles are released into

the bile from isolated perfused rat livers exposed to PAs (LaFranconi et al., 1985; Mattocks

et al., 1991; Yan and Huxtable, 1994). These pyrroles have been identified as 7-GSDHP

and its breakdown products (Lame etaL, 1990; Mattocks etal., 1991; Lame etal., 1995).

The presence of sulfur-conjugated pyrroles in the urine has also been demonstrated,

including a A^-acetyl-cysteine adduct (Estep et al., 1990; Mattocks and Jukes. 1992).

Conjugation widi glutathione by a pyrrolic metabolite of senecionine has been demonstrated

in vitro with hepatic microsome preparations (Buhler et al., 1990; Reed etal., 1992). It is

not known whedier GSH conjugation with DHAs is enzymatically or chemically mediated.

However, this reaction occurs rapidly in vitro without the presence of enzymes (Mattocks

and Jukes, 1990a, 1992b). Although a cysteine conjugate of DHP has not been identified,

its formation from the glutathione conjugate could be catalyzed by y-glutamyl

transpeptidase and dipeptidases localized in the kidney. Further metabolism of a cysteine

conjugate by (S-lyases could produce an unstable thiol which yields electrophilic products

(Vamvakas and Anders, 1990).

The role of 7-GSDEIP and the further putative metabolite 7-CYSDHP in

extrahepatic toxicity of PAs was a debated question for some years, with studies yielding

conflicting results (Huxtable et al., 1990; Cooper et al., 1992; Pan etal., 1993). It was

shown that 7-GSDHP and 7-CYSDHP can produce pneumotoxicity as measured by the

reduction of serotonin transport in the isolated perfused rat lung (LaFranconi et al., 1985)

Page 43: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

39

and produce pulmonary hyperplasia in rats in vivo (Huxtable et al., 1990), suggesting a

possible roie for these compounds in the pneumotoxicity of monocrotaline.

DHP

The pyrrolic metabolite DHP is formed by the hydrolysis of the DHAs. Hsu et al.

(1973) studied the distribution and toxicity of [^H]DHP, finding that the compound quickly

distributes throughout the body. The major accumulation and subsequent site of toxicity

was found in the glandular portion of the stomach. Mattocks (1986) suggested that this is

due to the acidity of this region, which causes an increase in the reactivity of this

compound. Thus local polymerization and alkylation by DHP is observed. DHP causes

pneumotoxicity (Huxtable et al., 1978) in rats, but doses of 100 mg/kg are needed,

compared to 60 mg/kg for monocrotaline. There are also qualitative differences in the

pneumotoxicity produced by DHP and monocrotaline, suggesting that DHP is not the sole

pneumotoxic metabolite. DHP has also been shown to be hepatotoxic, antimitotic,

immunosuppressive, teratogenic and carcinogenic (Peterson et al., 1972; Samuel and Jago,

1975; Peterson and Jago, 1980; Percy and Pierce, 1971; Peterson et al., 1983). Many of

the toxic effects of DHP are similar to that produced by the parent alkaloids, but higher

doses of the pyrrole are required. This is because DHP, a pyrrolic alcohol, is much less

reactive and more water-soluble than the corresponding DHA, an ester pyrrole. However,

because of the additional stability of DHP compared to the DHAs, it may travel to sites not

normally exposed to the DHAs. The lower reactivity and hence weaker alkylating activity

militate against DHP as the primary extrahepatic toxin, but it may be an important

component in the overall chronic toxicity expressed by a particular alkaloid.

Page 44: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

40

Effect of PA Structure on Metabolism and Toxicity

Tne cytotoxicity of PAs is associated witii their hepatic metabolism to pyrroie

derivatives. Several structural factors determine the extent of conversion of PAs to toxic

pyrroles in relation to non-toxic metabolites. Structures which impart resistance to ester

hydrolysis, increase lipophilicity and favor metabolic dehydration rather than A/^oxidation

tend to increase the relative toxicity of a PA (Mattocks, 1986). The relationship between

these factors effecting pyrrole production in an animal is very complex, and any one factor

is insufficient to explain the actions of a particular PA. Toxicity will also be affected by the

nature of the pyrrole metabolites produced. Primary pyrroles differ greatly in their stability,

which causes differences in their abilities to react with tissue nucleophiles, rates of

detoxication, or which organs they will be exposed to.

Lipophilicity

In general, the more lipophilic a PA, the more susceptible it is to bioactivation by

hepatic P-450s (Mattocks, 1981; Mattocks and Bird, 1983; Mattocks, 1986). Additionally,

a lipophilic DHA metabolite would be expected to cross the cell membrane of a target organ

easier than a more water soluble DHA. The hydrophilic PAs, such as the 9-monoesters

lycopsamine or indicine, are less extensively metabolized to pyrroles, are rapidly excreted,

and are weakly toxic. However, some highly lipophilic PAs or semi-synthetic derivatives

produce unexpectedly less pyrroles in animals because they are subject to extensive

enzymatic hydrolysis (Mattocks, 1981).

Ester hydrolysis

The hydrolysis of PAs by hepatic esterases is a detoxication pathway. PAs which

are normally innocuous become distinctly hepatotoxic in animals which have been treated

Page 45: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

41

with compounds (e.g. tri-orthocresylphosphate) to irreversibly inactivate esterase activity

(MaLlocks, 1981; Mallocks, 1986). This is because more PA becomes available for

bioactivation to pyrroles. Mattocks (1982) examined both the base and enzyme catalyzed

rates of hydrolysis of a series of PAs in vitro. Rates of hydrolysis under both conditions

were dependent on steric hindrance to the ester groups. Open ester PAs with simple,

unbranched acids were extremely sensitive to hydrolysis. Those with bulkier or branched

acid substitutions hydrolyzed much slower. Despite the rigidity of the ester groups,

macrocyclic PAs are less susceptible to hydrolysis than open type PAs. Again, those which

contain hindering groups near the ester portion are less prone to hydrolysis. The PA

jacobine contains an epoxide joining CI5 and C20 in the acid moiety, which is near the C7

ester. This substitution significantly inhibits hydrolysis by guinea pig carboxylesterases

compared to senecionine, which differs only in the lack of the epoxide function (Dueker et

al., 1995; Chung and Buhler, 1995). This may partially explain the increased susceptibility

of the guinea pig to jacobine, since more alkaloid would be available for bioactivation to

toxic pyrroles.

Bioactivation to DHAs versus N-oxidation

N^-oxidation represents a major detoxication pathway for PAs in many animals. The

balance between this pathway and dehydrogenation to pyrroles is important in the

expression of toxicity. Aside from lipophilicity, there are no clear-cut structural features

which are known to favor one pathway or another. Metabolic dehydrogenation is believed

to occur by an initial microsomal hydroxylation at C8 of the pyrrolizidine ring (Mattocks

and Bird, 1983) (Figure 5). For optimal pyrrole production, a minimum of steric hindrance

in the vicinity of C8 may be an important factor, although this would be expected to allow

increased access for hydrolysis. Macrocyclic diesters produce the greatest amounts of

Page 46: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

42

pyrrole metabolites compared to A/-oxides (Mattocks and Bird, 1983), in part due to their

high lipophiliciiy. Tne macrocyciic structure may favor pynole fomiation because tlie ring

is relatively rigid, being held away from the "top" of the necine, allowing access to C8 by

P-450 enzymes (Mattocks, 1986). The pyrrole to N-oxide ratio is highest in macrocyciic

diesters and open monoesters. This ratio is lowest from open diesters, which have the most

hindrance around C8 due to two flexible acid substitutions.

Little effect is observed between metabolism of heliotridine (hydroxyl at C7 is a)

based or retronecine (hydroxyl at C7 is |3) based semisynthetic PAs with simple ester

groups (Mattocks and Bird, 1983). This implies the conformation at C7 does not influence

microsomal metabolism. However, the natural PA heliotrine is consistently metabolized to

pyrrole to a greater extent than its isomer indicine both in vivo and in vitro, and is also

more toxic in vivo (Bull etal., 1958; Schoental, 1968; Mattocks, 1972). Thai the natural

heliotridine based PAs are frequently more toxic than retronecine based PAs may be due to

increased resistance to hydrolysis by the former (Mattocks, 1981).

DHA reactivity

While it is clear that PA structure affects the rate and extent of metabolism, structure

is also critical in determining the fate of the reactive metabolite once it has formed. These

compounds are highly reactive, with half-lives in water of a few seconds (Mattocks and

Jukes, 1990). The DHAs differ significantly with respect to alkylation reaction kinetics

(Karchesy and Deinzer, 1987) and rates of hydrolysis. Once formed, an extremely reactive

DHA may hydrolyze so fast that it does not have time to react with vital cellular

nucleophiles. If the metabolite has a slightly longer half-life, it may react with cellular

components in the vicinity of its formation, thus causing hepatotoxicity. A portion of

metabolites with longer half-lives may survive transport to organs downstream from the

Page 47: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

43

liver, particularly the lungs. Some DHA could conceivably survive long enough to cause

other symptoms such as neurotoxicity. On die other hand, if a DHA was sufficiently

unreactive, it may be excreted or largely dispersed before it can react with a macromolecule

(Mattocks, 1986).

In addition to stability, the number and posidon of reactive centers of the

dehydronecine moiety may influence toxicity of a PA. Diester PAs are believed to form

bifunctional alkyladng pyrroles, while the weakly toxic 9-monoester pyrroles only have a

single reactive center (Karchesy and Deinzer, 1987; Mattocks and Jukes, 1990a, 1990b).

The theory of bifunctionality is indirectly supported by studies which indicate that PA

metabolites cross-link DNA (Petry etai, 1984; Hincks etal., 1991). Additionally, only

macrocyclic diester PAs are known to produce pneumotoxicity. Driver and Mattocks

(1984) have demonstrated that synthetic derivatives of pyrrolic esters, "synthanecines", are

pneumotoxic only when diey posses two labile ester functions. Monoester synthanecines

(like monoester PAs) produced only hepatotoxicity. Whether this is attributable to cross-

linking of tissue nucleophiles by pyrroles is unknown.

Statement of the Research Problem and Hypothesis

PAs containing a I -2 double bond are hepatotoxic. Certain PAs, particularly some

of the macrocyclic diesters, produce toxicities in organs in addition to the liver. Of the PAs

used in this dissertation, monocrotaline is pneumotoxic, causing pulmonary hypertension

and right heart hypertrophy (Huxtable, 1990). Trichodesmine is a potent neurotoxin

(Shtenberg etal., 1955; Ismailov etai, 1970), while seneciphylline and retrorsine, except

under unusual circumstances (Mattocks and White, 1973), are hepatotoxic. The two open

ester PAs studied in this dissertation, lycopsamine and 7-acetyl-lycopsamine are known

solely as hepatotoxins. Bioactivation by the liver is a requirement for the expression of PA

toxicity. Mattocks (1968a) proposed that the primary metabolite of PAs were the

Page 48: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

44

corresponding DHAs; electrophilic compounds which, by covalently binding to important

tissue nucleophiies, were responsible for toxicity of the parent aikaioids. Subsequent work

has established that DHAs are indeed produced by the liver (Mattocks etal., 1990; Glowaz

et al., 1992). Why structurally similar PAs often produce markedly different toxicities is

poorly understood. Additionally, the nature of the extrahepatic metabolite(s) is still unclear.

Whether the DHAs are sufficiently stable to reach sites away from the liver has been an

ongoing question. The hypotheses of this dissertation are that (1) DHAs are responsible for

the extrahepatic toxicity observed with some PAs, (2) stmctural differences in the ester

portion of DHAs cause significant differences in their physicochemistry (e.g. lipophilicity,

stability) and (3) these physicochemical differences are partially responsible for differences

in patterns of their further metabolism and observed toxicity. To address these problems we

have investigated the in vivo toxicity of the recently discovered PA metabolite 7-GSDHP

and the further putative metabolite 7-CYSDHP, to determine their role as

extrahepatotoxins. The results of this study will provide support for or against the DHAs as

the proximate extrahepatic toxin. Secondly, to establish the comparative stability of a series

of structurally similar DHAs, we have examined their reactivity toward a variety of

nucleophiies. These studies include the potentiometric determination of DHA half-lives of

hydrolysis, rates of DHA alkylation by 4-(/?-nitrobenzyl)pyridine, a nitrogen nucleophile,

and their relative reactivity toward a large molecular weight sulfhydryl in the form of

Thiopropyl-Sepharose® resin. The relationship between relative stabilities of the DHAs

and metabolism and toxicity of PAs will be examined. We have also determined the

chemical structure of sulfur-bound pyrroles formed both in vitro and by the in vivo

metabolism of PAs. These experiments will determine the position of the reactive center of

the various DHAs, providing information on how structure affects the namre of alkylation.

To examine the effects of DHAs on a model cellular target, the inhibition by DHAs of yeast

and rat brain hexokinase was studied in vitro. Hexokinase, a critical enzyme in cellular

Page 49: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

45

energy production, is sulfhydryl rich and known to be inhibited by compounds which

alkylate these substitutions (Wilson, 1995). Finally, because monocrotaline is the only

commercially available PA, we have developed an efficient means of preparative separation

of PAs using high-speed counter-current chromatography. This has allowed us to utilize

locally and internationally collected plants as sources of structurally varied PAs. The

chromatographic technique enables the rapid separation and purification of quantities PAs

sufficient for these studies.

Page 50: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

46

METHODS

PA Isolation, Separation and Identincation

Plant material

Amsinckia tessellata A. Grey was collected near Tucson, Arizona, in the spring of

1994 and 1995. Senecio doiiglasii DC. var. longilobus (Benth.) L. Benson was collected

in the spring of 1995 from Sonoita, Arizona. Seeds from Trichodesma incaniim Alph. DC.

were collected from Uzebekistan in the summer of 1990. Symphytum spp., sold

commercially as comfrey root, was purchased from local supermarkets. Monocrotaline was

purchased commercially or isolated from seeds of Crotalaria spectabilis Roth, collected in

Georgia, and from C. retiisa L. collected in Natal, Brazil, in September of 1992.

Alkaloid isolation

PAs were isolated from plant material according to die method of Adams and

Rogers (1939). Prior to extraction, Symphytum root, and seeds from Crotalaria and

Trichodesma were finely ground in an electric coffee mill. Plant material was continuously

extracted in a Soxhlet apparatus with EtOH for 24 hr. Dried whole plant material from

Senecio and Amsinckia was ground in a commercial hammer mill through 3 mm mesh,

followed by soaking in EtOH for 4 days in a commercial extractor. All extracts were then

reduced under vacuum using a rotary evaporator. The residue was resuspended in HiO and

made acidic (pH 3-4) with citric acid and reduced with Zn dust overnight. The aqueous

solution was filtered, and extracted with equal volumes of CHCI3 until the organic phase

was clear. The aqueous portion was then made basic (pH 8-9) with 30% NH4OH and

extracted 3X with equal volumes of CHCI3. Aqueous phases from Amsinckia and

Symphytum extractions were saturated with NaCl to increase extraction of the polar

Page 51: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

47

alkaloids lycopsamine and intermedine (Culvenor er a/., 1980; Roitman, 1983). The

organic portions were combined, dried over anhydrous Na2S04, filtered and evaporated to

dryness, yielding the alkaloid fraction. Content of the alkaloid fractions was determined by

GC-electron impact MS (Method H) and TLC. Monocrotaline was obtained pure by

recrystallization from EtOH, as it was the only PA detected in the Crotalaria fractions.

Mixed PA fractions were separated and purified by counter-current chromatography as

described below. Alkaloids obtained (Figures 1 and 2), and their plant sources are listed in

Table 2.

High-speed counter-current chromatography

An Ito multilayer coil counter-current chromatograph equipped with a #10 coil

(PTFE tubing, 2.6 mm i.d., 380 ml capacity) was used for separations of alkaloidal

fractions isolated from Trichodesma, Senecio, Amsinckia and Symphytum. Potassium

phosphate buffer, 0.2 M, at an appropriate pH was used as the stationary phase. Mobile

phase was CHCI3, driven by a metering pump at a flow rate of 210 ml/hr in a head-to-tail

fashion. Prior to use, the two solvent phases were equilibrated by shaking together in a

separatory funnel, and then allowed to separate. The column was initially filled with

stationary phase, then rotated at 800 rpm while mobile phase was loaded. The column was

considered in equilibrium when only CHCI3 eluted from the tail end. At this point, alkaloid

samples, either solid or non-solid, 200-800 mg, were dissolved in 3-5 ml of 1:1 mobile

phaserstationary phase and injected into the column. Mobile phase was collected by a

fraction collector at 1.7 min intervals. Alkaloid remaining in the column after separation

was collected by purging the stationary phase with Ni, followed by basifying and CHCI3

extraction. Chromatograms were constructed by measuring relative PA content of each

fraction colorimetrically with Ehrlich reagent (2% boron trifluoride etherate, v/v and 2%

Page 52: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

48

/7-dimethylaminobenzaldehyde, w/v, in EtOH) on a spectrophotometer at 562 nm

(Maltocks, 1968). Fractions containing llie same alkaloid were combined, filtered, dried

under vacuum and if possible, recrystallized from EtOH or acetone. Corrected retention

time of solute peaks, r'R, was found by subtracting time of peak maximum of the solvent

front from time of peak maximum of the solute. Partition coefficients, K, defined as solute

concentration in the mobile phase divided by solute concentration in the stationary phase,

were estimated by the method of Zhang et al. (1988). All peaks were analyzed on TLC and

GC-MS (Method H) for identification and purity. Where electron impact-MS failed to

produce a molecular ion, flow injection-atmospheric pressure chemical ionization (APCI)

MS was used.

Thin-layer chromatography

PAs, pyrrolic metabolites and derivatives were analyzed by TLC as listed in Table

1. Plates spotted with PAs were developed by spraying with methanolic o-chloranil (0.5%

W/V), followed by Ehrlich reagent. Pyrroles were visualized by Ehrlich reagent.

Gas chromatography electron impact-mass spectrometry, method I

GC EI-MS was performed by the Mass Spectrometry Facility, Department of

Chemistry, University of Arizona, on a Hewlett Packard 5890 system using an HP-5 fused

silica capillary column consisting of 5% phenyl-methyl silicone, 25 m x 0.2 mm with He

carrier gas at 23 psig. Conditions: injector temperamre 250 °C; starting temperature

programmed 70°, held 1 min, to 300° at 20° min"', held 7 min; split inj. ratio 70:1, vol. I ftl

(CH2CI2). The column was connected directly to an HP 5970 mass selective detector. Mass

spectra were recorded at 70 eV.

Page 53: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

49

Gas chromatography electron impact-mass spectrometry, method II

Performed at the Southwest Environmental Health Sciences Center, University of

Arizona, on a Fisons MD800 using a DB-5, 15 m x 0.25 mm column with He carrier gas at

5 psig. Conditions: temperature program same as method I; spiitless injection volume 1 |il

(CH2CI2). The column was connected directly to a mass selective detector. Mass spectra

were recorded at 70 eV.

Flow-injection atmospheric pressure chemical ionization mass spectrometry

Flow-injection APCI MS was conducted on a Finnigan TSQ 7000 triple quadrupole

mass spectrometer at same location as GC-MS method H. Conditions: mobile phase MeOH

0.5 ml min ', samples introduced into the ion-source via direct inj., 50 |j.l inj. vol (MeOH),

10 eV offset added to the region between sample introduction and mass analysis to increase

fragmentation (in-source collision-induced decomposition). The TSQ was equipped with an

atmospheric chemical ionization source and was scanned 100 to 700 amu at I scan sec '.

Preparation of PA Metabolites and Derivatives

DMAs

DHAs were prepared from parent alkaloids according to Mattocks and Jukes

(1989). PAs (100 mg) were dissolved in CHCI3 (20 ml), added to a solution of o-bromanil

(125 mg) in CHCI3 (20 ml) and allowed to stand at room temperature for 2 min. This

solution was extracted with 10 ml of an ice-cold aqueous solution of sodium borohydride

(200 mg) and potassium hydroxide (5 g) by \'igorously shaking in a separatory funnel for

20 sec. The CHCI3 layer was quickly removed and dried over anhydrous Na2S04, filtered

and evaporated under vacuum with a rotary evaporator. TLC analysis indicated that parent

alkaloid or DHP was not present in the reaction product. DHAs were stored at -20 °C under

Page 54: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

50

nitrogen. DEIA samples were disgarded if polymerization was apparent, as indicated by the

presence of a reddish-brown color.

7-CYSDHP (7-cysteinyl-6,7-dihydro-1-hydroxymethyl-SH-pyrrolizine)

7-CYSDHP was synthesized as described by Mattocks and Jukes (1990).

Dehydromonocrotaline (100-500 mg) was dissolved in 1 ml of ethylene glycol dimethyl

ether, and added dropwise over 5 minutes to 4 ml of a stirred solution of L-cysteine

(equimolar to dehydromonocrotaline) in ammonium acetate buffer (1.0 M, pH 7.5). The

reaction was carried out on ice, under N2and monitored by a pH meter. A 10% solution of

NH4OH was used to maintain a constant pH of 7.5-7.6 during the course of mixing by

adding dropwise when the pH of the reaction mixture began to fall. The resultant solution

was allowed to remain on ice for an additional 10 min. Precipitates were removed by

centrifiigation. The supernatant was lyophilized, leaving a pinkish-white solid material.

Positive ion FAB-MS showed an (M+H)"^ ion, m/z 257 (Figure 42). TLC indicated the

presence of 2 Ehrlich positive spots, which were also ninhydrin positive. FAB-MS was

conducted by the Midwest Center for Mass Spectrometry, University of Nebraska, Linoln,

Nebraska.

7-GSDHP (7-glutathionyl-6,7-dihydro-1-hydroxymethyl-SH-pyrrolizine)

7-GSDHP was synthesized in an analogous manner to 7-CYSDHP. Positive ion

FAB-MS indicated an (M+H)"^ ion, m/z 443.

DHP ((±)-6,7-dihydro-7-hydroxy-1-hydroxymethyl-SH-pyrrolizine).

DHP, a racemic mixture of dehydroretronecine and dehydroheliotridine, was

prepared by the method of Mattocks and Jukes (1989).

Page 55: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

51

9-Ethoxy DHP (7-hydroxy-9-ethoxy-6,7-dihydro-5H-pyrrolizine)

Ethoxy pyrroles were synthesized according to Mattocks and Jukes (1990a, 1992).

A solution of dehydrolycopsamineZ-intermedine (10 mg), was allowed to stand overnight in

5 ml of EtOH containing 0.2 ml of 1 M NaOH. Aqueous K2CO3 (5% w/v, 2 ml) was

added and mixed for 5 min. The ethanol was removed by reduced pressure. The aqueous

portion was then extracted with 3X with CHCI3 and dried over anhydrous Na2S04. The

combined organic extracts were combined, filtered and dried under vacuum to yield 9-

ethoxy DHP. Structures of all pyrrolic ethyl ethers were confirmed by TLC and GC-MS

(Method I).

7-Ethoxy DHP ((±)-7-ethoxy-1-hydroxymethyl-6,7-dihydro-5H-pyrrolizine)

Dehydromonocrotaline (5-10 mg) was dissolved in 5 ml of EtOH containing 0.2 ml

of 1 M NaOH. Isolation of the products proceeded as above. Product was primarily

7-ethoxy DEff. with a small amount of the diether present.

7,9-Diethoxy DHP ((±)-7-ethoxy-9-ethoxy-6,7-dihydro-5H-pyrrolizine)

The diethoxy pyrrole was produced as above using dehydromonocrotaline (5-10

mg) as the starting material and unmodified ethanol (5 ml) as the solvent. Workup and

isolation was identical to the monoethers.

Toxicity of Sulfur-Conjugated Metabolites of PAs, in vivo

7-CYSDHP

Rats were obtained from the University of Arizona Division of Animal Resources

and were housed four to a cage in 12 hr cycles of light and dark. They were allowed food

and tap water ad libitum. Rats (60-70 g) were randomly divided into three groups of six

Page 56: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

52

animals. 7-CYSDHP, dissolved in normal saline, was administered via a single

intravenous tail vein injection (0.1 - 0.3 nil), at 300 and 600 p.niol/k.g. Control rats were

injected with normal saline. Body weight and overall health of the animals were recorded

daily. At 21 days, rats were weighed, killed and necropsied. Liver, lung, heart, thymus,

spleen and kidneys were removed for determination of tissue to body weight ratios. Wet

weights were immediately recorded, after which organs were dried at 50 °C for 48 hr prior

to obtaining dry weight. In order to detect evidence of pneumotoxicity and subsequent right

heart hypertrophy, the right heart to left heart plus septum weight ratio (RV/LV + S) was

determined. This was performed by dissecting the right atrium free from the left heart and

septum prior to weighing. Differences in means between treatment groups was determined

by ANOVA with the aid of a computer program (Pharm-Calc; PCS, Phildelphia, PA).

Group means were considered statistically different at /? < 0.05.

7-GSDHP

Rats weighing 160-190 g were randomly divided into three groups of six animals.

Groups were injected via tail vein with either 7-GSDHP in saline at 600 |i,moi/kg,

monocrotaline at 185 (imol/kg as a positive control, or saline. Animals and data collection

were treated as above for the remainder of the study.

Chemistry of Sulfur-Bound Pyrroles

Preparation and activation ofThiopropyl-Sepharose 6B resin

Desiccated resin (2 g) was suspended in 100 ml potassium phosphate buffer (pH

12, 0.2 M) and allowed to swell for 60 min with occasional stirring. The suspension was

centrifuged, the buffer decanted and the procedure repeated until resin had been washed

with 500 ml of buffer. When the buffer was decanted a final time, the resin had swelled to

Page 57: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

53

a total volume of 6 ml. Buffer was added (14 ml) so that each ml of suspension contained

100 mg di7 weight of resin. Tlie resin was activated by adding 5 of p-mercaptoellianoi Lo

resin aliquots (2 ml suspension, containing 200 mg resin dry wt, approximately 22 nmol

thiol groups), and allowing them to stand for 2 hr at room temperature. Prior to use, the

resin was rinsed with buffer in a small glass column connected to a vacuum. The resin was

rinsed until |3-mercaptoethanol could not be detected (by the absence of yellow color) in the

eluent using Ellman's Reagent (1% 5,5'-dithiobis(2-nitrobenzoic acid) in acetone). Fully

activated resin contains approximately 105 ^mol of suifhydryls per g dry wt.

Recirculating aqueous flow system

Covalent binding of DHAs to immobilized thiols was studied using an aqueous

flow system described by Mattocks and Juices (1990b) (Figure 7). This system was

designed to mimic the slow release of DHA into the circulation from the liver. A small

quantity of DHA dissolved in a non-protic solvent was continuously injected by a syringe

pump over 1 hr into a fast moving stream of buffer (potassium phosphate, pH 7.2, 37 °C.

7.5 ml/min), rapidly mixed, and was passed over a bed of activated Thiopropyl-Sephorose

resin (200 mg dry wt containing 20 |j.mol of sulfhydryl groups) contained in a small glass

column. The buffer was recirculated from a reservoir via a peristaltic pump. The amount of

time between DHA injection and exposure to the resin was determined by the length of

tubing (polyethylene, 1 mm i.d.) through which the mixture must travel prior to passing

through the resin bed. This was determined by recording the amount of time for a small air

bubble to pass through the length of tubing. Dye was injected into the system via the

syringe pump to visually confirm that mixing was instantaneous and complete. At the

conclusion of each run, the resin was examined for quantity of bound pyrrole or the

position of sulfur binding as described below.

Page 58: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

54

Identification ofThiopropyl-Sepharose resin-bound pyrroles

Thiopropyl-Sepharose resin was exposed to various DHAs using the recirculating

system described above. Upon injection, each DHA (10 i^mol) was exposed to the buffer

for either 1.5 or 30 sec prior to passing through the resin (5 and 400 cm tubing,

respectively). Following the 1 hr run, the resin was rinsed with water, then with EtOH

sufficient to remove all water. The resin was removed from the column and suspended in 5

ml of EtOH and stirred. An aqueous solution of Tris-buffered silver nitrate (1 ml) was

added dropwise (silver nitrate (400 mg) was mixed with aqueous (2 ml) tris-

(hydroxymethyl)mediylamine (560 mg) and neutralized with dilute HNO3 (Mattocks and

Jukes (1990a)). The solution was allowed to stir for 30 min, followed by addition of 10%

aqueous potassium carbonate (2 ml). The solution was allowed to stir for an additional 5

min followed by centrifugation to pellet solids. The supernatant was retained and the

ethanolic portion removed under reduced pressure. The remaining aqueous portion was

brought to 5 ml with HiO, then extracted with CHCI3. The organic layer was dried over

anhydrous Na2S04 and evaporated under reduced pressure. The resultant pyrrolic ethyl

ether was identified by comparing with standards on TLC and by GC-MS (Method I)-

Identification of liver sulfiir-bound pyrroles

Pyrrole metabolites covalently bound to liver and bile of PA treated rats were

released in the form of soluble pyrrolic ethyl ethers (Mattocks and Jukes, 1990). Male

Sprague-Dawley rats (120-130 g) were given a single i.p. injection of either monocrotaline

(3(X) |imol/kg), retrorsine (75 |imol/kg), seneciphylline (150 |imol/kg), trichodesmine (50

nmol/kg) or lycopsamine/intermedine (500 ^mol/kg). Animals were sacrificed after 24 hr.

Livers were removed and portions (5 g) were homogenized in EtOH and centrifuged 5 min

at 2700 rpm. Pellets were resuspended in EtOH (20 ml) and rapidly stirred. Aqueous Tris-

Page 59: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

55

buffered silver nitrate (4 ml) was added dropwise and stirring continued for 30 min.

Workup of Ihe homogenale proceeded identically as with the resin. Pyrrolic ethyl ethers

were identified by TLC and GC-MS.

Bile samples (200-500 ̂ 1), collected from isolated rat livers perfused with PAs

from earlier experiments (Yan et al., 1995), were immediately frozen. Thawed bile samples

were treated with Tris-buffered ethanolic silver nitrate in an analogous fashion to liver

homogenates and resin.

Quantitation ofDHA binding to Thiopropyl-Sepharose resin

Relative pyrrole binding to activated Thiopropyl-Sepharose resin was measured

following its exposure to various DHAs using the recirculating aqueous flow system

described above. The amount of time between DHA mixing and exposure to the resin was

kept as short as possible (1.5 sec; 5 cm tubing) to minimize loss of DHA to hydrolysis.

DHAs (6 |j.mol in I ml dimethylformamide) were injected over one hour. Following each

run, bound pyrrole was quantified according to Mattocks and Jukes (1990b). The resin

(while still in the column) was rinsed with HiO followed by MeOH. The resin was

removed and suspended in 5 ml of MeOH. Aliquots (3) of suspension (1 ml) were

removed and I ml of Ehrlich reagent was added to each, followed by heating in a water

bath at 60 °C for I min. After the solution had cooled, 1 ml of 10% methanolic HgCK was

added and the solution was vortexed for 30 sec. The solution was brought to 5 ml with

MeOH and centrifuged to pellet the resin. The supernatant was read at 562 nm in a

spectrophotometer. Pyrrole content was determined by using absorbance 8.7 per |imol of

pyrrole released (as read in 5 ml) (Mattocks and Jukes, 1990b).

Page 60: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

56

Studies of DHA Reactivity

Alkylaliun of DHAs by 4-(p-nilrobenzyl)pyridine

AJkylation reactions of DHAs by 4-(/7-nitrobenzyl)pyridine were performed as

described previously (Karchesy et ai, 1987). Briefly, 250 nmol of DHA, dissolved in 25

Hl dimethylformamide was added to 250 ^1 of 5% 4-{/7-nitrobenzyl)pyridine in acetone

acidified with 25 nl of formic acid. Reactions were carried out at room temperature (23°C).

Blanks contained no DHA. The reaction was quenched by the addition of ice cold

acetone:triethylamine, 1:1. The resultant blue-purple color was read spectrophotometrically

at 565 nm. Samples were diluted with ice-cold acetone as necessary. Results for each DHA

were plotted as total absorbance versus time. To determine rate constants, ^i, derived

curves were fit to a monoexponential equation (1), or a biexponential equation (2) in the

case of dehydroretrorsine, with the aid of a computer program (Karchesy et ai, 1987).

(1) A = A»-A«„e'*'

(2) A = A„-Be"*'-Ce"*'

Potentiometric measurement of hydrolysis of DHAs

Hydrolysis rates for a series of DHAs was determined potentiometrically by

recording the fall in pH on hydrolysis due to release of necic acid. DHAs (10 |imol

dissolved in 100 |al dimethylformamide) were rapidly added to a stirred HEPES solution

(10 ml, 0.5 mM, pH 8). The change in pH was continuously followed by a pH meter

connected to a chart recorder, previously calibrated from pH 3 to 10. A sigmoidal log-linear

plot was produced which was linear through at least two pH units. From this slope, the rate

constant, k, and the half-life of hydrolysis, txa could be estimated by assuming single

component exponential decay. Specifically, k was found by using the concentrations of the

base form of HEPES (pATa 7.55) at pH 6 and 7, calculated from the Henderson-

Page 61: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

57

Hasselbalch equation. These values were then used to derive k from an equation describing

single component exponential decay, x = Ae ", which describes the linear portion of the

generated curves. Time r, measured directly from the chart recordings, is the time to go

from pH 7 to 6, A is the concentration of HEPES base at time 0 (pH 7), and x is the

concentration of HEPES base at time t (pH 6). Values derived represent the disappearance

of HEPES conjugate base, which was assumed to be directly proportional to appearance of

hydrogen ion liberated from ester hydrolysis. Half-lives were found by using the equation

0.691k. Data was analyzed for statistical differences by Newman-Keuls multiple analysis

with the aid of a computer program (Tallarida and Murray, 1986).

Inhibition of Hexokinase, in vitro

Yeast hexokinase assay

Yeast hexokinase (E.G. 2.7.1.1., from Baker's yeast) activity was measured in a

coupled assay with glucose-6-phosphate dehydrogenase by following the reduction of P-

NADP spectrophotometrically (Bennet er iz/., 1962; Puri e/a/., 1988). Assay volumes

were 1 ml and contained the following: triethanolamine (TEA) 39 mM; D-glucose, 0.01 -

216 mM; ATP, 0.74 mM; MgCU, 7.8 mM; P-NADP, 1.1 mM; hexokinase, 0.01 units:

glucose-6-phosphate dehydrogenase, 1.0 unit, pH 7.6. Mixtures were allowed to

equilibrate in quartz cuvettes (1 cm light padi) at 25 °C for 5 min prior to addition of

hexokinase. Reactions were read at 340 nm every 30 sec for 7 min. Final enzyme activity

was measured as change in absorbance per min over the 7 min assay. The enzyme rates

were linear over the course of measurement. For inhibition studies, hexokinase (0.75 units

in 960 III of 100 mM TEA buffer, pH 7.6, 25 °C) was incubated for 2 min with either

dehydrotrichodesmine, dehydroretrorsine, dehydro-7-acetyl-lycopsamine or

dehydrolycopsamine (dissolved in 40 ^1 dimethylformamide) at 0.01 - 20 mM. An aliquot

Page 62: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

58

(15 nl, representing 0.01 units of hexokinase) of these incubations was then added to the

above described assay solution and residual phosphotransferase activity was measured.

Vehicle controls lacked DHA. DHP controls were performed by adding DHAs to the

incubation mixture 1 min prior to adding the hexokinase. Similar studies were performed

with mannose (10 mM) and ATP (10 mM) present in the incubation mixture to determine

whether they afforded protection to hexokinase against DHA inhibition. Data are plotted as

residual hexokinase activity versus concentration of DHA. Studies were also performed to

determine hexokinase activity in the presence of glucose concentrations from 0.01 to 5.0

mM. Prior to assaying, the enzyme was exposed to either vehicle, or dehydrotrichodesmine

at 2.5, 5.0 and 10 mM in the manner described above. Data were plotted as hexokinase rate

(nmol NADPH/min) versus concentration of glucose. and data were derived from

non-linear regression with the aid of a computer program (SigmaPlot, Jandel Scientific,

San Raphael, Ca) Data were also plotted as [S]/y versus [S] in order to graphically

illustrate non-competitive inhibition.

Rat brain hexokinase assay

Rat brain hexokinase was assayed according to Bennet et al. (1962). Male Sprague-

Dawley rats were used, weighing between 150-200 g. Animals were killed by decapitation,

the brain immediately removed and weighed. Brains were then homogenized in a Teflon-

glass homogenizer in 10 ml of ice cold potassium phosphate buffer (100 mM, pH 8.0) per

gram of tissue. Homogenates were then diluted with buffer to a concentration of 5 mg wet

tissue/ml and stored on ice. Assay solution (I ml) contained glycylglycine, 150 mM;

MgCU, 20 mM; ATP, 9 mM; P-NADP, 1 mM; and glucose, 25 mM. Assays were

conducted in quartz cuvettes at 37 °C. Solutions were prewarmed for 5 min prior to the

addition of 50 |al of brain homogenate (representing 0.25 mg wet tissue) and glucose-6-

Page 63: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

59

phosphate dehydrogenase, 0.3 units. For inhibition studies, 960 |il of brain homogenate

was incubated wilii DHAs (2 min) or PAs (10 aiin), dissolved in 40 |i.l dinietliylforniaiuide

at 1.0 - 10 mM. Vehicle controls lacked DHA. Studies were also conducted in which

DHAs were added to the incubation buffer (910 ̂ ll) 5 min prior to adding a concentrated

aliquot of brain homogenate (50 [il undiluted homogenate). Assays were read every 30 sec

for 10 min at 340 nm using a spectrophotometer.

Glucose-6-phosphate dehydrogenase assay

Glucose-6-phosphate dehydrogenase (EC 1.1.1.49; from Torula yeast) was

assayed according to Noltman etal. (1961). In a 1 ml assay solution, concentrations were

glycylglycine, 50 mM, pH 7.4; D-glucose 6-phosphate, 2 mM; |3-NADP, 0.67 mM;

MgCli, 10 mM; glucose-6-phosphate dehydrogenase, 0.05 units. For inhibition studies,

glucose-6-phosphate dehydrogenase was prepared at 1 unit/ml in glycylglycine buffer, 50

mM, pH 7.4. Studies with DHAs were conducted in the same fashion as with yeast

hexokinase. Assays were conducted at 25 °C, read every 30 sec for 7 min at 340 nm.

Page 64: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

60

Table 1

TLC Condilions for PAs aiid Pyrrolic Derivatives

PA or Pyrrole Plate Solvent Rf

Monocrotaline silica a 0.53

Trichodesmine C18 b 0.57

Incanine C18 b 0.42

Retrorsine silica c 0.35

Senecionine silica c 0.61

Seneciphylline silica c 0.63

Florosenine silica c 0.97

Lycopsamine* silica c 0.22

3-Acetyl-lycopsamine* silica c 0.46

7-Acetyl-lycopsamine* silica c 0.66

3,7-Diacetyl-lycopsamine* silica c 0.92

Symphytine* silica c 0.84

7-CYSDHP CIS d 0.74, 0.26

7-GSDHP silica e 0.54, 0.19', 0.86'

DHP CI8 f 0.83

7-Ethoxy DHP CI8 f 0.69

9-Ethoxy DHP C18 f 0.56

7,9-Diethoxy DHP CI8 f 0.31

^•BuOH : EtOH : H,0 : NH3OH (30%), 24 : 14 : II : 1 "EtOH : H.0 : NH;OH(30%), 60 : 39; I ^CHClj: MeOH : NH,OH (30%), 85 : 14 : I '' EtOH : H,0 : NH3OH, 89 : 10 : 1

MeOH : CHCI3 : H.O : EtjN. 80 : 8.5 : 11 : 0.5 f EtOH : H,0 : NH3OH, 50 : 49 : I. * One or more stereoisomers present, with identical Rf values ' trace

Page 65: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

61

syringe pump

mixer

trapping column

delay loop

pump sepharose resin

37 "bath reservoir

Circulation Apparatus

Figure 7. Schematic of recirculating aqueous flow system for "trapping" of DHAs.

Page 66: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

62

RESULTS

High-Speed Counter-Current Chromatography

High-speed counter-current chromatography was used for the separation of PA

mixtures isolated from Amsinckia tessellata, Symphytum spp. (comfrey root),

Trichodesma incamim, and Senecio douglasii var. longilobus (Figures I and 2). A two-

component solvent system was chosen, consisting of a chloroform mobile phase and

potassium phosphate buffer, 0.2 M, at various pH values (Table 2). as the stationary

phase. Up to 800 mg of sample could be separated in a single run, with excellent resolution

of alkaloids. However, the diastereomers lycopsamine and intermedine, or their

derivatives, could not be resolved.

A. tessellata collected near Tucson, Az., contained 0.02% PA per dry weight of

plant. These PA mixtures were isolated as a brown gum, as reported previously (Culvenor

and Smith; 1966 Roitman, 1983). Figure 8 shows the resolution of the PAs using an

aqueous phase pH of 7.4, from 710 mg of this gum. Three alkaloid peaks were obtained

and additional PA remained in the column. Peaks I-EH, plus the retained PA, yielded a

single spot on TLC. However, analysis by GC-MS indicated each peak consisted of an

isomeric pair, producing nearly identical mass spectra, but with slightly different retention

times. PAs were of both open mono- and diesters of retronecine. Peak I, 3,7-diacetyl-

lycopsamineZ-intermedine (a mixture of the two isomers 3,7-diacetyl-lycopsamine and 3,7-

diacetyl-intermedine) (5.0 mg) was retained sufficiently to elute after the colored impurities,

and was well resolved from peak H, 7-acetyl-lycopsamine/-intermedine (13.1 mg). Peak

EH, 3-acetyl-lycopsamine/-intermedine (55.1 mg), completely eluted by 2 hr, but the peak

exhibited some tailing. Lycopsamine/intermedine (70 mg) remained in the column and were

recovered by purging. Mass spectra were consistent with literature data (Kelley and Seiber,

Page 67: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

63

1992). As reported by other researchers (Culvenor and Smith, 1966; Roitman, 1983), the

alkaloids, which were obtained as amber oils, could not be crystallized.

Several samples of commercial comfrey root from different sources were extracted,

with alkaloid content ranging from 0.01-0.09% dry weight. The PA fraction was isolated

as a brown gum, consisting of open mono- and diesters. Figure 9 shows the resolution of

586 mg of alkaloidal fraction from comfrey, using a stationary phase pH of 5.6. Two well

resolved peaks were obtained, with a third alkaloid remaining in the column. The colored

impurities, unretained by the stationary phase, eluted with the solvent front. TLC indicated

one spot for each eluted peak and the retained material, although GC-MS revealed the

presence of isomers. Peak I (19.0 mg) consisted of 3 isomeric PAs, all producing nearly

identical mass spectra, consistent with that reported for symphytine and its 3 known

diastereomers (Kelley and Seiber, 1992). Peak II, 7-acetyl-lycopsamine/-intermedine (13.6

mg), was retained until 113.1 min, due to the low pH of the stationary phase compared to

the Amsinckia separation. Lycopsamine and intermedine (210.6 mg) remained in the

column. Following chromatography, PAs were obtained as amber oils, but again could not

be crystallized.

S. douglasii var. longilobus collected from Sonoita, Az, contained ca. 0.03% PAs

per dry weight of plant material, isolated as a clear-brown gum. GC-MS analysis of the

alkaloidal fraction revealed the presence of the CI 2 macrocyclic type PAs senecionine,

seneciphylline, retrorsine, and the otonecine-based florosenine, previously unreported in S.

douglasii. The PAs had mass spectrum matching literature spectrum (Cava et al., 1968;

Quails and Segall, 1978). Riddelliine, reported to occur in S. douglasii var. longilobus,

was not detected (Adams and Govindachari, 1949). Figure 10 shows a chromatogram of

the separation of 600 mg of this alkaloidal fraction. Excellent resolution was achieved at a

stationary phase pH of 5.0. Peak I, florosenine (2.0 mg), was unretained at this or any

other pH however, and eluted together with the colored material at the solvent front.

Page 68: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

64

Table 2

Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography

Stationary t' * ' R

Species Alkaloid phase pH" (min) K"

Amsinckia 3,7-Diacetyl-lycopsamine/-intermedine 7.4 2.6 27.4

tessellata 7-Acetyl-lycopsamine/-intermedine 13.1 6.4

3-Acetyl-lycopsamine/-intermedine 74.0 1.3

Lycopsamine/Intermedine''

Symphytum Symphytine and/or isomers 5.6 14.5 6.4

spp. 7-Acetyl-lycopsamine/-intermedine

Lycopsamine/Intermedine''

113.1 0.9

Trichodesma Incanine 6.0 7.7 11.3

incaman Trichodesmine 36.6 2.6

Senecio Florosenine 5.0 0.9 52.0

douglasii

var. Senecionine 34.9 2.7

longilobus Seneciphylline

Retrorsine''

70.6 1.4

"Solvent system composed of CHCI3 mobile phase and phosphate buffer, 0.2 M, at indicated pH, as the stationary phase.

'detention time, defined as time of solute peak maximum minus time of solvent front. Tartition coefficients, K, estimated according to Zhang etal. (1988). ''Alkaloid(s) remained in colunui and were recovered as described in the Methods section.

Page 69: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

65

m

s a

CN vO »o

(U CJ c rt jD

LA o zn

<

0

h

0 20 40 60 80 100 120 140

Time (min)

Figure 8. Separation of PAs from Amsinckia tessellata using high-speed counter-current chromatography. I = 3,7-Diacetyl-lycopsamine/-intermedine; II = 7-Acetyl-lycopsamine/-intermedine; EI = 3-Acetyl-lycopsamine/-intermedine. LycopsamineZ-intermedine was recovered from the column as described in Methods.

Page 70: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

0.8 s c

CN vo irj © 0.6

0.2

0.0

80 100 120 140 160 180 0 20 40 60

Time (min)

Figure 9. Separation of PAs from Symphytum spp. using high-speed counter-current chromatography. I = Symphytine and isomers (symlandine, myoscorpine, echiumine); II 7-Acetyl-lycopsamine/-intermedine. LycopsamineZ-intermedine was recovered from the column as described in Methods.

Page 71: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

67

0.4 r

0.3

0.2

0.1

0.0 w

J. J-

0 20 40 60 80

Time (min)

100 120 140

Figure 10. Separation of alkaloids from Senecio douglasii var. longilobus using high-speed counter-current chromatography. I = Florosenine; 11 = Senecionine; HI = Seneciphylline. Retrorsine remained in the column and was recovered as described in Methods.

Page 72: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

Figure 11. Separation of PAs from Trichodesma incanum using high-speed counter-current chromatography. I = Incanine; II = Trichodesmine.

Page 73: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

69

Peak n, senecionine (23.0 mg), and peak IE, seneciphylline (63.7 mg) were well resolved,

while relrorsine (216.6 mg) remained in the column, and was easily recovered after the

run. With the exception of florosenine, alkaloids were recrystallized from acetone. TLC

and GC-MS analysis of peaks revealed single alkaloids, although TLC failed to separate the

structurally similar senecionine and seneciphylline (Chizzola, 1994; Cooper et al., 1996).

Alkaloidal fraction from seeds and surrounding capsule of T. incanum yielded

1.1% PA by weight, isolated as white crystals. In agreement with an earlier account

(Yunusov and Plekhanova, 1957), the mixture of alkaloids consisted of two cyclic diesters

of retronecine, incanine and trichodesmine, as determined by GC-MS. Figure 11 shows the

chromatogram of a separation of 250 mg of the alkaloid fraction, using a stationary phase

pH of 6.0. Two well resolved peaks were obtained; peak I, incanine (7.3 mg) and peak II,

trichodesmine (238 mg). Separations were performed with up to 800 mg of alkaloid, but in

higher amounts the trichodesmine peak tailed considerably, although resolution was still

satisfactory. Although peak U exhibited slight shoulders, no other alkaloids were detected

from this peak by GC-MS. TLC and GC-MS analysis of each peak revealed the presence

of a single PA, confirming complete resolution. The mass spectrum of incanine, which has

not been previously reported, is shown in Figure 29.

Toxicity of Sulfur-Conjugated Pyrroles, in vivo

The role of sulfur-conjugated pyrroles as potential toxic metabolites of PAs was

investigated by examining the in vivo effects of the PA metabolite 7-GSDHP, and the

further putative metabolite, 7-CYSDHP. Animals were killed at 21 days because

pulmonary hypertension and cor pulmonale associated with a single dose of monocrotaline

is of delayed onset (Huxtable. 1990).

Page 74: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

70

7-CYSDHP

7-CYSDHP was administered to male rats (60 - 80 g) via single tail vein injection at

3(X) and 600 ^mol/kg. Animals were necropsied, and organ and body weights (Tables 3-5)

was recorded as described in Methods. Over the course of the study, animals in the high

dose group gained significantly less weight than controls, although they appeared healthy.

Both dose groups had significantly reduced wet and dry liver to body weight ratios.

Pneumotoxicity was apparent by the presence of a dose-related increase in lung mass. Both

groups displayed significantly increased wet lung to dry lung weight ratios in comparison

to control animals. The high dose group also had significantly elevated wet lung/body

weight ratio. An increase in right heart size was observed in both treatment groups, but

values were not significantly different, possibly due to the large sample deviation.

7-GSDHP

7-GSDHP was administered to male rats (160 - 190 g) via single tail vein injection

at 600 p.mol/kg. Monocrotaline was given as a positive control at 185 |amol/kg

(approximately half the acute LDjg value). Rats treated with monocrotaline displayed

characteristic hepato- and pneumotoxicity (Tables 6-8). Significant increases in both wet

and dry organ/body weight ratios were observed in liver, lung, spleen, right ventricle and

RV/LV+S compared to control and 7-GSDHP treated animals. Contrary to some earlier

studies (Huxtable et al., 1990; Bowers et al., 1991), rats treated with 7-GSDHP showed

little signs of toxicity throughout the study. A mild increase in right heart size occurred, but

was not significant. The right heart/left heart plus septum ratios were significantly elevated

however, but less so than the monocrotaline treated group. An increase in this ratio

indicates right heart hypertrophy. However, this may be due to the correspondingly low

Page 75: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

71

Table 3

Summary of Wet Organ Weight / Body Weight Ratios in Rats at 21 Days Following a Single Injection (i.v.) of 7-CYSDHF'

Control CYSDHP CYSDHP

Organ (Saline) (300 ^mol/kg) (600 M-mol/kg)

Liver 5.002 ± 0.285 4.614 ±0.095* 4.402 ± 0.242''

Lung 0.559 ± 0.057 0.648 ± 0.088 0.694 ± 0.084*

Spleen 0.312 ±0.027 0.336 ±0.017 0.324 ± 0.025

Thymus 0.277 ± 0.031 0.321 ±0.028 0.324 ± 0.049

Kidney 0.852 ± 0.052 0.876 ± 0.036 0.860 ± 0.020

Right Ventricle 0.069 ± 0.005 0.070 ±0.011 0.075 ± 0.008

Left Ventricle 0.272 ± 0.008 0.287 ±0.010 0.282 ±0.013

RV/LV+S^ 0.251 ±0.025 0.244 ±0.031 0.266 ± 0.029

Weight Gain (g) 146.1 ±4.596 139.7 ±11.25 132.3 ± 9.728*

"Organ weights expressed as means (x 10') ± SD (N = 6). ''Significantly different than controls at p < 0.05. "^V/LV+S = right ventricle weight/ left ventricle plus septum weight.

Page 76: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

72

Table 4

Summary of Dry Organ Weight / Body Weight Ratios in Rats at 21 Days Following a Single Injection (i.v.) of 7-CYSDHP

Control CYSDHP CYSDHP

Organ (Saline) (300 ^lmol/kg) (600 nmol/kg)

Liver 1.559 ±0.058 1.377 ±0.038'' 1.322 ±0.102"

Lung 0.115 ±0.016 0.127 ±0.017* 0.136 ±0.018"

Spleen 0.068 ± 0.006 0.073 ± 0.005 0.070 ± 0.005

Thymus 0.123 ±0.015 0.132 ±0.008 0.133 ±0.015

Kidney 0.213 ±0.013 0.219 ±0.009 0.215 ±0.005

Right Ventricle 0.014 ±0.001 0.014 ±0.002 0.015 ±0.002

Left Ventricle 0.059 ±0.001 0.062 ± 0.002 0.060 ± 0.003

RV/LV+S' 0.231 ±0.018 0.222 ± 0.025 0.247 ± 0.028

"Organ weights expressed as means (x 10") ± SD (N = 6). ^Significantly different than controls zip < 0.05. '"RV/LV+S = right ventricle / left ventricle plus septum ratio.

Page 77: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

73

Table 5

Summary of Wet Organ Weight / Dry Organ Weight Ratios in Rats at 21 Days Following a Single Injection (i.v.) of 7-CYSDHP

Control CYSDHP CYSDHP

Organ (saline) (300 |imol/kg) (6(X) |imol/kg)

Liver 3.205 ± 0.085 3.352 ±0.102 3.332 ± 0.093

Lung 4.873 ± 0.073 5.084 ±0.071'' 5.113 ±0.115"

Spleen 4.609 ± 0.099 4.591 ±0.077 4.604 ± 0.095

Thymus 4.944 ±0.091 4.963 ±0.103 4.896 ± 0.096

Kidney 4.395 ±0.147 4.472 ±0.186 4.119 ±0.067

Right Ventricle 4.961 ±0.125 5.055 ±0.313 5.026 ±0.215

Left Ventricle 4.634 ± 0.075 4.596 ± 0.054 4.673 ± 0.068

"Data expressed as means ± SD (N = 6). ''Significantly different than controls at p < 0.05.

Page 78: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

74

Table 6

Summary of Wet Organ / Body Weight Ratios in Rats at 21 Days Following a Single Injection (i.v.) of 7-GSDHP or Monocrotaline"

Control Monocrotaline GSDHP

Organ (saline) (185 nmol/kg) (600 ^lmol/kg)

Liver 3.687 ± 0.292 4.319 ± 0.363" 3.996 ±0.197

Lung 0.494 ± 0.024 0.819 ±0.076" 0.497 ± 0.034

Spleen 0.279 ± 0.048 0.354 ± 0.029'' 0.260 ± 0.024

Thymus 0.250 ± 0.042 0.251 ±0.026 0.243 ± 0.020

Kidney 0.837 ± 0.032 0.858 ± 0.025 0.807 ± 0.052

Right Ventricle 0.067 ±0.014 0.089 ± 0.004" 0.075 ± 0.008

Left Ventricle 0.317 ±0.045 0.287 ±0.013 0.278 ±0.017

RV/LV+S'' 0.212 ±0.020 0.311 ±0.016" 0.268 ± 0.029'"

Weight Gain (g) 239.8 ±24.1 239.8 ± 7.4 253.0 ± 12.0

"Organ weight data expressed as means (x 10") ± SD (N = 5, except 7-GSDHP, N = 6). ''Significantly different from control and 7-GSDHP treated at p < 0.05. "Significantly different from control and monocrotaline treated at p < 0.05. ''RV/LV-t-S = right ventricle weight / left ventricle plus septum weight.

Page 79: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

75

Table 7

Summary of Dry Organ / Body Weight Ratios at in Rats at 21 Days Following a Single Injection (i.v.) of 7-GSDHP or Monocrotaline"

Control Monocrotaline GSDHP

Organ (Saline) (185 limol/kg) (600 lamol/kg)

Liver 1.120 ±0.093 1.320 ± 0.064'' 1.29 ±0.080"

Lung O.I 11 ±0.011 0.162 ±0.011'' 0.105 ±0.006

Spleen 0.064 ± 0.009 0.081 ± 0.007'' 0.060 ± 0.005

Thymus 0.067 ± 0.002 0.055 ± 0.005" 0.053 ± 0.004

Kidney 0.201 ±0.012 0.207 ± 0.004 0.193 ±0.014

Right Ventricle 0.015 ±0.003 0.019 ±0.001" 0.017 ±0.002

Left Ventricle 0.067 ± 0.004 0.064 ± 0.003 0.064 ± 0.003

RV/LV-hS'' 0.220 ± 0.033 0.302 ± 0.020" 0.265 ± 0.025"

"Organ weight data expressed as means (x 10") ± SD (N = 5, except 7-GSDHP, N = 6). ''Significantly different from control and 7-GSDHP treated at p < 0.05. "Significantly different from control and monocrotaline treated at;? < 0.05. ''RV/LV-f-S = right ventricle weight/ left ventricle plus septum weight.

Page 80: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

76

Table 8

Summary of Wet Organ / Dry Organ Weight Ratios in Rats at 21 Days Following a Single Injection (i.v.) of 7-GSDHP or Monocrotaline"

Control Monocrotaline GSDHP

Organ (Saline) (185 |imol/kg) (600 [xmol/kg)

Liver 3.122 ±0.088 3.263 ±0.212 3.082 ± 0.773

Lung 4.478 ± 0.253 5.045 ±0.174* 4.728 ±0.129

Spleen 4.363 ±0.140 4.395 ±0.189 4.321 ±0.078

Thymus 4.603 ± 0.050 4.600 ±0.051 4.622 ± 0.054

Kidney 4.175 ±0.121 4.159 ±0.189 4.321 ±0.052

Right Ventricle 4.467 ±0.179 4.654 ±0.177 4.399 ±0.131

Left Ventricle 4.748 ± 0.549 4.521 ±0.132 4.345 ±0.112

T>ata expressed as means (x 10") ± SD (N = 5, except 7-GSDI-IP, N = 6). ''Significantly different from controls and 7-GSDHP treated at /y < 0.05.

Page 81: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

77

weights of the left heart from this group, which would artificially increase the ratio. A small

increase in iiver weight was observed, but was not significantly different than controls.

Chemistry of Sulfur-Bound Pyrroles

Identification ofThiopropyl-Sepaharose resin-bound pyrrole

Sulfur-bound pyrrole can be solvolized by silver or mercury ion (Mattocks and

Jukes, 1990a) with cleavage of the thioether bond. If this procedure is conducted in a

buffered ethanolic solvent, the former thioether will be replaced with an ethoxy

substitution. The pyrrole can be identified on TLC or GC-MS, establishing the position of

the ethoxy group (C7 or C9), and thus the position of sulfur-binding. We have used this

procedure with Thiopropyl-Sepharose resin exposed to different DHAs. Each DHA (10

Hmol in 1 ml dimethylformamide) was continuously injected into the recirculating system

over a period of one hour. Delay times were 1.5 and 30 sec (5 cm and 400 cm tubing

length, respectively). Resin exposed to dehydrotrichodesmine, dehydromonocrotaline,

dehydroseneciphylline and dehydroretrorosine, and dehydro-7-acetyl-lycopsamine yielded

7-ethoxy DHP, regardless of the delay time. Thus, time of exposure to the buffer (1.5 or

30 sec) did not affect the pattern of alkylation. Evidence of bialkylation was not observed

as 7,9-ethoxy DHP was not detected. Pyrrole released from resin exposed to

dehydrolycopsamine after a 1.5 sec hydrolysis time was identified as 9-ethoxy DHP. This

indicates that all DFIAs except dehydrolycopsamine bound the resin at C7 of the pyrrole

moiety. Dehydrolycopsamine bound the resin at C9. When dehydrolycopsamine was

allowed a 30 sec exposure to the buffer prior to trapping, no pyrrole could be detected from

the resin.

Page 82: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

78

Chemistry of liver sulfur-bound pyrroles

Bile was collected from isolated rat livers perfused with monocrotaline,

seneciphylline, retrorsine, trichodesmine and a mixture of lycopsamine and intermedine

(Yan et at., 1995). Rats were given single i.p. doses of the same alkaloids at 65-90% acute

LDjo (3-5 day) levels (Mattocks, 1986; Wild, 1994) and killed at 24 hr. Liver homogenates

from rats injected with PAs, and bile from PA perfused livers were treated with Tris-

buffered ethanolic silver nitrate to determine the position of sulfur-binding of pyrrole

metabolites. When bound pyrroles were cleaved from either liver or bile sulfhydryls, 7-

ethoxy DHP was produced, as determined by TLC and GC-MS (Method I). This indicated

the presence of C7 pyrrole conjugates from liver and bile. As with the resin, no evidence of

sulfhydryl crosslinking was detected, as a pyrrolic diether was not produced in the silver

nitrate reactions. Pyrrolic ethers were not detected in bile or tissue from rat liver perfused

with the C9 monoesters lycopsamine and intermedine. Thus a C9 sulfur adduct did not

form, or was present in very small amounts. Lung and brain from the PA treated animals

were treated in the same manner as the livers, but pyrrolic ethers were not detected. This

was attributed to the fact that tissues were assayed under buffered conditions, which

decreases the sensitivity of this procedure (Mattocks and Jukes, 1990a).

Quantitation ofDHA binding to Thiopropyl-Sepharose resin

The relative reactivity of a series of DHAs toward a large molecular weight

sulfhydryl compound was studied by using Thiopropyl-Sepharose as a model nucleophile.

Equimolar amounts of DHAs (6 nmol) were injected into the recirculating aqueous flow

system, where they were instantaneously mixed with the buffer and passed through the

activated resin after a 1.5 sec delay (5 cm tubing). The resin-bound pyrrole was cleaved by

ethanolic silver nitrate and quantified with Ehrlich Reagent. Table 9 shows the amount of

Page 83: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

79

Table 9

Relative Quantity of Pyrrole Bound to Thiopropyl-Sepharose Resin After Exposure to DHAs"

Bound pyrrole

DHA (^imol) {% of total injected)

Dehydromonocrotaline 0.80 ±0.01 13.31 ±0.26

Dehydroseneciphylline 0.76 ± 0.05 12.71 ± 0.81

Dehydro-7-acetyl-Iycopsamine 0.58 ±0.08 9.59 ± 1.39

Dehydroretrorsine 0.51 ± 0.05 8.50 ± 0.76

Dehydrotrichodesmine 0.42 ±0.04 7.07 ± 0.62

Dehydrolycopsamine* 0.05 ± 0.02 0.85 ± 0.24

"Data from recirculating system as described in Methods. DHAs (6 ^mol) were exposed to the aqueous environment for approximately 1.5 sec prior to passing through the resin. Values are means ± S.D. (N = 4-9 experiments). different from other groups at p < 0.01.

Page 84: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

80

pyrrole subsequently bound to the resin after exposure to different DHAs. The highest

quantity of pyrrole was trapped on passing dehydromonocrotaline and

dehydroseneciphylline through the resin, with 0.80 ^mol and 0.76 nmol trapped,

representing 13.3 % and 12.7%, respectively, of total DHA injected. Pyrrole trapped from

dehydro-7-acetyl-lycopsamine, dehydroretrorsine and dehydrotrichodesmine was 0.57

M-mol, 0.51 ^imol, and 0.42 ^lmol, representing 9.6% , 8.5%, and 7.1% respectively, of

total injected DHA. A mixture of the 9-monoesters, dehydrolycopsamine and

dehydrointerraedine, was found to react minimally with the resin, with 0.09 [imol trapped

pyrrole, or 0.85% of the total injected. DHP, 7-GSDHP, and parent PAs did not react with

the resin in this system.

Hydrolysis of DHAs

When a DHA is introduced into an aqueous environment, the ester rapidly

hydrolyzes forming DHP and the corresponding necic acid. As a result, the pH of the

solution falls. DHAs, prepared according to established methods (Mattocks et al., 1989),

were rapidly added to a stirred HEPES solution (0.5 mM, pH 8.0). Change in pH was

followed potentiometrically, producing a sigmoidal log-linear curve which was near linear

through at least 2 pH units (Figure 12). From the slope of the near linear portion, the rate

constant, k, and the half-life, of hydrolysis under the studied conditions could be

estimated. Calculations were based on the disappearance of HEPES conjugate base, which

was assumed to be directly proportional to liberation of protons by ester hydrolysis.

Relative rates of hydrolysis for the DHAs are shown in Table lO.Under the conditions

studied, the two open esters, dehydrolycopsamine and 7-acetyl-lycopsamine, had the

shortest half-lives; 1.02 and 0.39 sec, respectively. Dehydroretrorsine and

Page 85: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

81

9

8

7

6

4

3 20 35 40 5 10 15 25 30 0

Time (sec)

Figure 12. Representative potentionietric recording of the hydrolysis of dehydroretrorsine and dehydrotrichodesmine. The fall in pH was due to the release of protons on hydrolysis.

Page 86: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

82

Table 10

Rate Constants and Half-Lives of Hydrolysis of DHAs"

DHA )t(sec ')±S.D.' S.D/

Dehydro-7-acetyl-lycopsamine 1.77 ±0.09 0.39 ± 0.02

Dehydrolycopsamine 0.68 ±0.12 1.02 ±0.22

Dehydroretrorsine 0.65 ±0.10 1.06 ±0.20

Dehydroseneciphylline 0.43 ± 0.06 1.60 ±0.28

Dehydromonocrotaline 0.20 ± 0.06 3.39 ± 0.99

Dehydrotrichodesmine 0.13 ±0.02 5.36 ± 0.92

"Values were determined potentiometrically as described in Methods. Data represent means ± SD from a minimum 6 experiments.

''All values were statistically different at p < 0.01 except for dehydrolycopsamine and dehydroretrorsine, whose values did not differ significantly.

Tlalf-lives calculated by Q.69/k

Page 87: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

83

dehydroseneciphylline were slightly more stable, with half-lives of 1.06 and 1.60 sec,

respectively. Dehydromonocrotaline and dehydrotrichodesmine were the slowest to

hydrolyze, with half-lives of 3.39 and 5.36 sec, respectively. All half-life values were

significantly different at /? < 0.01, except for dehydroretrorsine and dehydrolycopsamine,

which did not differ.

Alkylation of 4-(p-nitrobenzyl)pyridine by DHAs

The DHAs reacted with 4-(p-nitrobenzyl)pyridine under pseudo-first-order

conditions (Karchesy and Deinzer, 1987), (Table 11) forming an adduct which could be

detected colorimetrically after quenching with base at 565 nm (Figures 13, 14). As time

increased, the amount of detected color increased, indicating buildup of product. The two

C12 DHAs reacted at the highest rates with 4-(p-nitrobenzyl)pyridine. Dehydroretrorsine

was the most reactive, the data fitting a biexponential rate expression (/:,, = 0.226 min ',

= 0.021 min"'). Dehydroseneciphylline, another C12 macrocycle, reacted about 4.5 times

slower (fc, = 0.041 min ') with the nucleophile. The faster reaction, of dehydroretrorsine

and 4-(p-nitrobenzyl)pyridine appeared to be mostly over by 20 min, at which time the

slower, second color-forming reaction, could be observed. Dehydrotrichodesmine (A:, =

0.013 min"'), a CI 1 DHA, was the slowest reacting compound tested.

Dehydromonocrotaline (fc, = 0.026 min"') reacted at twice the rate of

dehydrotrichodesmine. The alkylation of dehydroseneciphylline, dehydromonocrotaline

and dehydrotrichodesmine fit monoexponential rate expressions. The open ester DHAs

were also reactive toward 4-(/?-nitrobenzyl)pyridine. The diester dehydro-7-acetyl-

lycopsamine reacted twice as fast as its 9-monoester derivative, dehydrolycopsamine, with

values of 0.152 and 0.076 min"', respectively.

Page 88: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

Table 11

Rate Constants for the Alkylation of DHAs by 4-(/?-nitrobenzyl)pyridine''

DHA A:, (min ') ± S.E. r''

Dehydroretrorsine 0.226 ± 0.063 0.996

Dehydro-7-acetyl-lycopsamine 0.152 ±0.011 0.980

Dehydrolycopsamine 0.076 ± 0.003 0.994

Dehydroseneciphylline 0.041 ±0.002 0.999

Dehydromonocrotaline 0.026 ± 0.002 0.993

Dehydrotrichodesmine 0.013 ±0.001 0.994

"Reactions were carried out under pseudo-first order conditions at 22 °C. Rate constants calculated from mono- or biexponential rate equation. 'Tlegression coefficient.

Page 89: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

85

• Dehydroretrorsine

A DehydroseneciphylHne

• Dehydromonocrotaline

• Dehydrotrichodesmine

0 -

10 20 30 40 50 60

Time (min)

Figure 13. Rates of alkylation of macrocyclic DHAs by 4-05-nitrobenzyl)pyridine. Data are means ± standard deviations from 3 observations. Data for dehydroretrosine was fit to a biexponential rate expression. Data for the remaining DHAs were fit to a monoexponential equation.

Page 90: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

86

• Dehydro-7-acetyl-lycopsamine

• Dehydrolycopsamine

c c ca vo

<u (J c

X3 U« O C/3

X) <

0 0 10 20 30 40 50 60

Time (min)

Figure 14. Rates of alkylation of open-ester DHAs by 4-(/7-nitrobenzyl)pyridine. Data are means ± standard deviations from 3 observations. Data were fit to a monoexponential rate equation.

Page 91: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

87

Inhibition of Hexokinase by DHAs, in vitro

Yeasl hexokinase

The phosphotransferase activity of yeast hexokinase was irreversibly inhibited by

dehydrotrichodesmine in a dose-dependent fashion (Figure 15). Inhibition of the enzyme

occurred at concentrations of DHA above 1 mM and complete, or near complete

inactivation of hexokinase occurred at 20 mM dehydrotrichodesmine. Concentrations of

DHAs above 20 mM were not tested because of excessive polymerization in the incubation

mixture by the pyrrole. Dehydroretrosine and dehydro-7-acetyl-lycopsamine, DHAs widi a

7-ester, both inhibited the enzyme in a similar fashion to dehydrotrichodesmine. The 9-

monoester DHA, dehydrolycopsamine, was a weaker inhibitor of hexokinase than the

diester DHAs (Figure 16). Under the smdied conditions, the IC50S for

dehydrotrichodesmine and dehydrolycopsamine were 5.3 and 52.4 mM, respectively. IC50

values for other 7-ester DHAs did not differ from dehydrotrichodesmine (Data not shown).

The inhibitory action of the DHAs was lost if they were added to the incubation mixture 1

min prior to adding hexokinase, indicating that DHP and necic acid did not affect enzyme

activity under the studied conditions. Hexokinase samples incubated 30 or 60 min with

DHAs showed no recovery of activity. At 20 mM, the parent alkaloids trichodesmine and

retrorsine had no effect on hexokinase activity.

In order to assess the nature of enzyme inhibition by DHAs in terms of competitive

or non-competitive, hexokinase inhibition assays were conducted in the presence of

increasing concentrations of substrate. Figure 17 indicates that dehydrotrichodesmine

produced a dose-dependent decrease in the of hexokinase. Additionally, because a

rightward shift in the concentration-effect curves was not observed, the data suggest a non­

competitive interaction between enzyme and inhibitor. Non-linear regression analysis of the

data indicate a dose-related decrease in while values are similar (Table 12). When

Page 92: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

88

the data were plotted as [substrate]/!^ versus [substrate], the regressions were also

suggestive of a non-competitve interaction (Figure 18).

Experiments were also conducted in which hexokinase was preincubated with

mannose (10 mM) or ATP (10 niM) prior to dehydrotrichodesmine exposure. Mannose, a

high affinity substrate for hexokinase, provided the enzyme minimal protection against

inactivation by dehydrotrichodesmine (Figure 19). However, ATP, a co-substrate,

provided even greater protection to hexokinase against inactivation by

dehydrotrichodesmine (Table 13). From this experiment, die IC50 value for

dehydrotrichodesmine was 5.84 mM. When hexokinase was pretreated with mannose or

ATP, the IC50 values for dehydrotrichodesmine were 7.14 and 23.99 mM, respectively.

Rat brain hexokinase

Preliminary investigations were performed using rat brain hexokinase to determine

if inhibition occurred toward the mammalian enzyme. Rat brain hexokinase, assayed in

vitro, was significantly inhibited by 1 min incubation with 5 mM macrocyclic DHAs

(Figure 20) compared to controls. The enzyme was only slighdy inhibited by the DMF

vehicle (Figure 20). No effect on hexokinase was observed with dehydrolycopsamine at 5

mM or 10 mM (Figures 20, 21). Preincubation of brain homogenates with 1 mM glucose

afforded slight, but significant protection to hexokinase against 10 mM

dehydrotrichodesmine (Figure 21). Dehydrotrichodesmine was significantly more active

against hexokinase when the brain homogenates were preincubated with

P-mercaptoethanol (0.1%) for 24 hr prior to assay (Figure 21). p-Mercaptoethanol has a

reducing effect on several of the hexokinase disulfides (Lenzen et ai, 1989). As with yeast

hexokinase, inhibitory action of DHAs was lost if the compounds were added to the

Page 93: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

89

incubation buffer I min prior to brain homogenate. Additionally, parent PAs had no

inhibitory effect on enzyme activity.

Glucose-6-phosphate dehydrogenase

In comparative assays, dehydrotrichodesmine, at 1.0, 5.0 or 20 mM, did not

significantly inhibit glucose-6-phosphate dehydrogenase in relation to controls (data not

shown).

Page 94: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

90

Table 12

Effect of Dehydrotrichodesmine on Yeast Hexokinase Phosphotransferase Activity in the Presence of Increasing Substrate Concentrations"

[Dehydrotrichodesmine]

(mM)

V max

(nmol NADPH/min) (mM glucose) r'

0.0 9.76 ± 0.64 0.122 ±0.034 0.988

2.5 9.12 ± 1.28 0.165 ±0.089 0.969

5.0 4.80 ± 0.05 0.148 ±0.007 0.999

10.0 1.76 ±0.06 0.168 ±0.001 0.992

"Values were derived from non-linear regression of data from Figure 17 with the aid of a computer program (SigmaPlot®, Jandel Scientific, San Raphael, CA). Data represent means ± standard errors from 5 observations. '' Coefficient of regression.

Table 13

Effect of Mannose (10 mM) and ATP (10 mM) Pretreatment on the Inhibition of Yeast Hexokinase by Dehydrotrichodesmine"

Treatment ICjg

Dehydrotrichodesmine 5.84 ±1.06

Dehydrotrichodesmine + Mannose 7.14 ± 1.03

Dehydrotrichodesmine + ATP 23.99 ±1.05

"Data derived from regression analysis from linear portion of log concentration-effect curves (Figure 19) with the aid of a computer program (Tallarida and Murray, 1986). Data represent means ± standard errors from 5 observations. All data are statistically different at p < 0.05.

Page 95: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

[Dehydrotrichodesmine] (mM)

Figure 15. Concentration-effect curve of the Inhibition of yeast hexokinase by dehydrotrichodesmine. Data represent means ± standard deviations from 5 observations.

Page 96: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

100

80

60

Dehydrotrichodesmine Dehydrolycopsamine

40

20

t I t U 5 10

[DHA] (mM)

20

Figure 16. Concentration-effect curve for the inhibition of yeast hexokinase by dehydrotrichodesmine and dehydrolycopsamine. Data represent means ± standard deviations from 5 observations.

Page 97: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

93

10.0

8.0

6.0

4.0

2.0

• 0.0 mM Dehydrotrichodesmine

T 2.5 mM Dehydrotrichodesmine

• 5.0 mM Dehydrotrichodesmine

• lO.O mM Dehydrotrichodesmine

0.0

-U- • • • '

0.01 O.l 10

[Glucose]

Figure 17. Saturation curves for the inhibition of yeast hexokinase by dehydrotrichodesmine. Curves represent logistic non-linear regression fit with the aid of a computer program. Data represent means ± standard deviations from 5 observations.

Page 98: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

94

500

400

300

200

100

0 9 3 5 4

[Substrate] mM

Figure 18. [Substrate]A^ versus [Substrate] plot of the inhibition of yeast hexokinase by dehydrotrichodesmine. Data, transformed from Figure 17, represent means from 5 observations.

Page 99: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

95

120 • Dehydrotrichodesmine

• Dehydrotrichodesmine + Mannose (10 mM) • Dehydrotrichodesmine + ATP (10 mM)

100

80

60

£

'•w U < (U c« ca c • o X (U

± 40 C3 3

u

^ 20

0 .J L

10 20

[Dehydrotrichodesmine] (mM)

Figure 19. Effect of ATP and mannose on the inhibition of yeast hexokinase by dehydrotrichodesmine. Data represent means ± standard deviations from 5 observations. Hexokinase was preincubated with ATP or mannose for 5 min prior to exposure to dehydrotrichodesmine.

Page 100: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

96

• Control

OH Vehicle Control

Dehydrotrichodesmine (5 mM)

Dehydromonocrotaline (5 mM)

Hi Dehydroseneciphylline (5 mM)

Dehydroretrorsine (5 mM)

V/Wi Dehydrolycopsamine (5 mM)

Figure 20. Effect of DHAs on rat brain hexoidnase, in vitro. Data represent means ± standard deviations from 5 observations. *Indicates group mean is different than control at p < 0.05.

Page 101: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

97

160

140

120

100

80

- 60

40

20

0

• Vehicle Control

mnH Dehydrolycopsamine (10 mM)

Dehydrotrichodesmine (10 mM)

Dehydrotrichodesmine (10 mM) + Glucose Pretreatment (1 mM)

EZ3 P-MEtOH Pretreatment (0.1 %) Control

P-MEtOH Pretreatment (0.1 %) + Dehydrotrichodesmine (10 mM)

Figure 21. Effect of dehydrotrichodesmine and dehydrolycopsamine on rat brain hexokinase, in vitro. Hexokinase was preincubated 24 hr with P-mercaptoethanol and 5 min with glucose prior to dehydrotrichodesmine exposure. Data represent means ± standard deviations from 5 observations. *Indicates group means are different from control at p < 0.05. **Indicates group means are different from control and dehydrotrichodesmine at p < 0.05.

Page 102: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

98

DISCUSSION

PAs are found in a wide variety of flowering plant families distributed throughout

the world. They are toxic to animals, including man. Humans ingest PAs through the use

of herbal products or by the consumption of contaminated grains (Huxtable, 1989).

Hepatotoxicity is the trademark of PA poisoning, although other organs may be affected in

certain instances. In humans, ingestion of PAs results in veno-occlusive disease of the liver

and neurotoxicity in some cases (Ismailov etal., 1970). In addition to hepatotoxicity, PAs

such as monocrotaline produce pulmonary toxicity and associated cor pulmonale in the rat.

Hepatic metabolism of PAs by cytochrome P450s to pyrroles is a requirement for toxicity.

Metabolism by other organs has not been demonstrated, suggesting that the liver is the sole

source of toxic metabolites. Toxicity is believed to result from the formation of pyrrolic

ester intermediates (DHA), which are highly reactive electrophiles. Covalent binding of

DHAs to important tissue nucleophiles presumably initiates the events leading to toxicity.

In recent years, much information has emerged regarding the metabolism of PAs. It is now

widely accepted that DHAs are the proximate toxin, responsible for both hepatic and

extrahepatic toxicity. However, many important questions remain unanswered. Perhaps

most interesting is the observation that structurally similar PAs often produce markedly

different toxicities, in terms of both lethality and organs affected. The Senecio alkaloids,

such as seneciphylline and retrorsine are hepatotoxic in rats, except under unusual

circumstances (White et al., 1972; Mattocks and White, 1973; Ohtsubo et ai, 1977), in

which pneumotoxicity has been reported. The acute LD50S (3-5 day) of retrorsine and

seneciphylline in male rats are 97 and 230 [imol/kg, respectively (Mattocks, 1972; Bull et

ai, 1968). Crotalaria alkaloids such as monocrotaline are pneumotoxic in rats. The LDjg of

monocrotaline is 335 p.mol/kg (Mattocks, 1972). Trichodesmine is neurotoxic in humans,

rats and mice (Ismailov, 1970), yet differs from monocrotaline only by an isopropyl

Page 103: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

99

substitution at C14 rather than a methyl. Trichodesmine is also approximately six-fold more

lethal than monocrotaline, having an LD50 of 57 |imol/kg in rats. The open ester PAs are

only known to produce hepatotoxicity. The 9-monoesters are weak hepatotoxins in

comparison to macrocyclic diester PAs. Indicine, a stereoisomer of lycopsamine, has a

LD50 of greater dian 3.3 mmol/kg in rats (Schoental, 1968). The toxicity of open diester

PAs appears to vary greatly, although they are often as toxic as macrocyclic diesters.

Symphytine, a diester alkaloid of comfrey root, has an of 340 |imol/kg in rats (Hirono

et al., 1979). The reasons for these differences in toxicity are poorly understood.

We believe that the differences in PA toxicity are a reflection of differences in rates

of formation of DHAs and their further metabolism. We posmiate that the metabolic

disposition of DHAs is partially determined by differences in their stability, especially at the

7-ester position. The aim of this dissertation was to investigate the relationship between

physicochemical properties of structurally varied DHAs and observed toxicity of the

corresponding parent PAs. The goals of this research were to: 1) to develop an efficient

chromatographic system for the separation and purification of large quantities of PAs, 2)

investigate the role of sulfur-conjugated pyrroles as potential toxic metabolites of PAs, 3)

to use in vitro and in vivo methods to investigate differences in reactivity and stability of

structurally similar DHAs, 4) to examine the relationship between the nucleophilic

reactivity of these DHAs and patterns of their further metabolism in the isolated perfused rat

liver, and to the toxicity produced by the parent PA.

Counter-Current Chromatography

As plants typically contain several alkaloids, there has been a need for convenient,

rapid separation and purification methods for these compounds. The separation of PAs on

newer high speed counter-current chromatography instruments has not been reported until

recently (Cooper et al., 1996), and had been little used for alkaloids in general compared to

Page 104: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

100

other natural products of plant origin (Marston and Hostettman, 1994). We demonstrate the

preparative separation and purification of PAs isolated ixom Amsinckia tesseiiata,

Symphytum spp. (comfrey root), Trichodesma incanum, and Senecio douglasii var.

longilobus, using high speed CCC.

A chloroform (mobile) and phosphate buffer (stationary) solvent system was

chosen for separations of PAs. The large density difference between the two solvents

promotes rapid phase separation and excellent stationary phase retention even at high

mobile phase flow rates (Marston and Hostettman, 1994). PAs are easily recoverable from

a chloroform mobile phase. With an aqueous buffer stationary phase, good solute

resolution can be obtained since structurally similar PAs differ in their values

(Mattocks, 1986). Additionally, the colored impurities often associated with PA isolations

are largely unretained in an aqueous stationary phase, and elute at the solvent front.

Resolution of PAs is determined by partitioning between the two liquid phases. Partition

coefficients, K, defined as concentration of solute in the mobile phase divided by solute

concentration in the stationary phase, ideally fall between 2 and 0.5 for the most efficient

separations (Conway, 1990), but a wide range of K values may be acceptable depending

on the amount and retention times of impurities (Zhang et al., 1988). Partitioning of PAs

between solvent phases depends on the proportion of alkaloid in the ionized form, as well

as its lipid solubility in the free base form. Partition coefficients and retention time, of

PAs will therefore be profoundly affected by the stationary phase pH. Because the

stationary phase is a liquid, retained PAs are easily recovered, and their adsorptive loss to

various solid chromatographic media as reported by Zalkow etal. (1988), is prevented.

Amsinckia tessellata (Boraginaceae) is a member of a small genus of around 14

species native to North and South America and introduced to Australia (Culvenor and

Smith, 1966; Kelley and Seiber, 1992). Alkaloid content of Amsinckia is variable, but all

contain hepatotoxic PAs, dominated by the diastereomers lycopsamine and intermedine,

Page 105: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

l O I

with smaller quantities of their acetylated derivatives and other uncommon PAs. These are

all open mono- and diesters of retxonecine and heliotridine, 28 of which have been

identified (Kelley and Seiber, 1992). The abundance of A. tessellata in the vicinity of

Tucson, Arizona, makes this plant an ideal source for the open ester PAs used in this

dissertation. Roitman (1983) used silica gel column chromatography for preparative

separation of PAs from A. menziesii, while Culvenor and Smith (1966) employed counter-

current distribution to resolve PAs from several Amsinckia species. Both these techniques

proved time-consuming since successive runs were necessary for complete resolution. In

the genus Amsinckia, the relative quantities of stereoisomers is variable between both

species and different populations of the same species (Kelley and Seiber, 1992). Because

of the difficulty in resolving isomeric PAs on a preparative scale, comparative toxicological

studies for such stereoisomers have not been performed. Human poisonings by Amsinckia

have not been reported, but they are implicated in livestock poisonings in California

(Roitman, 1983).

Comfrey, consisting of dried root or leaves of Symphytum officinale (common

comfrey) or 5. x uplandicum Nyman (Russian comfrey, Boraginaceae), is widely used by

humans as a herb, vegetable and food supplement (Huxtable, 1990). Symphytum spp.

usually contain a complex mixmre of PAs, consisting of open mono- and diesters of

retronecine. They are mildly hepatotoxic and are carcinogenic in rats (Hirono et al., 1978),

and have been implicated in human poisonings (Huxtable, 1990; Ridker et al., 1985).

Culvenor et al. (1980) used counter-current distribution with a solvent system of

chloroform and phosphate buffer, pH 7.5, to resolve 1.8 g of an alkaloid fraction from S.

X uplandicum, yielding eight PAs. Due to the expense of comfrey root in the local

supermarkets, the variable PA content and difficulty in botanical identification, Amsinckia

was the preferred source for the open-ester PAs.

Page 106: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

102

The large genus Senecio (Asteraceae) has over 100 PA-containing species

(Huxtable, 1989). Senecio doiigiasii var. iongiiobus (Asteraceae), contains the hepatotoxic

PAs riddelliine, senecionine, retrorsine and seneciphylline (Adams and Govindachari,

1949), all structurally similar cyclic diesters of retronecine. In the southwestern U.S., S.

douglasii has been used extensively for ethno-medicinal purposes (Huxtable, 1990). Infant

poisonings and deaths have occurred in Arizona due to consumption of herbal preparations

containing S. douglasii var. Iongiiobus (Stillman etal., 1977; Fox etal., 1978). Zalkow et

al. (1988) using DCCC with a CHClj-CgHg-MeOH-HjO solvent system, separated a 3 g

mixture of 10 alkaloids from S. anonymus, including partial resolution of the diastereomers

senecionine and integerrimine. This separation demonstrates the resolving power of

counter-current techniques, although DCCC separations take longer, and choice of solvent

systems are limited (Conway, 1990). Resolution of the PAs senecionine and seneciphylline

in preparative systems has been difficult. Separation was achieved on silica gel

chromatography only by the use of sequential runs (Adams and Govindachari, 1949). The

two PAs were unresolvable by TLC in these and other studies (Chizzola, 1994). Despite

these earlier difficulties, they were well resolved in the present CCC system. Rorosenine

has been reported from Cacalia floridana, Senecio aureus, S. floridanus, S. fluviatilis

(Mattocks, 1986), but not from S. douglasii var. Iongiiobus (Adams and Govindachari,

1949). The former Senecio species are not known to occur in Arizona (Kearney and

Peebles, 1960), therefore contamination by another florosenine containing plant was

unlikely.

The small genus Trichodesma (Boraginaceae), native to Africa and Asia, contains

about 35 species, 3 of which are known to contain toxic PAs (Huxtable, 1989). T.

incanum has been responsible for outbreaks of human and livestock poisonings, reviewed

in the Russian literature (Smimov etal., 1959a, 1959b; Ismailov, 1970; Ismailov etal.,

1970). T. incanum has been shown to contain two neurotoxic PAs, incanine and

Page 107: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

103

trichodesmine (Yunusov and Plekhanova, 1957), both cyclic diesters of retronecine.

Previously, Irichoclesinine isolated from various Crotalaria was separated from other PAs

by column chromatography using Celite and Florosil, preparative TLC and DCCC (Adams

and Gianturco, 1956; Rothschild et al., 1979; Roeder et al., 1992). In the present study,

incanine and trichodesmine were co-isolated as white crystals. The lack of non-alkaloid

contaminants allowed for rapid separation times (< 2 hr), since the PA did not have to be

retained longer than impurities which elute near the solvent front. Thus, the high K value

for incanine (Table 2) was acceptable.

The use of high speed CCC in the preparative separation and purification of natural

products has become well established (Marston and Hostettman, 1994). Our results show

the value of this method for PAs. Plant compounds, including PAs, which are often

isolated as a gum or tar, as in the case of Senecio, Amsinckia and Symphytum alkaloid

fractions in this study, can be safely injected onto the column without fear of contaminating

or fouling a solid matrix. Thus pre-separation clean-up steps are unnecessary. Further

advantages demonstrated in this investigation include rapid separation times and low

solvent consumption (< 2 1 per run). The simple chloroform-phosphate buffer solvent

system used in these experiments was sufficient for the separations performed, although

the resolving power was insufficient to separate diastereomers. Preliminary experiments

indicate that ternary solvent systems of CHClj-MeOH-HjO may also be useful in

separating PAs. More elaborate solvents and techniques will become necessary with the

separation of increasingly complex PA mixtures. New techniques such as pH-zone-refining

CCC described by Ma etal. (1994), which utilizes a pH gradient, should prove especially

useful in cases where PAs possess large differences in K values.

Page 108: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

104

Toxicity of 7-CYSDHP and 7-GSDHP

Glutatliione is aii iniporlaiil cellular component in the defense against damage

produced by electrophilic compounds. Glutathione is found in high concentrations in many

mammalian cell types, participating in a variety of reactions. The nucleophilic sulfur atom

of the cysteine residue serves to neutralize many electrophilic metabolites. Conjugation is

generally mediated by the enzyme glutathione 5-transferase, although some highly

electrophilic compounds can react non-enzymatically (Koob and Dekant, 1991). Depending

on its disposition, a glutathione 5-conjugate may be further metabolized to a mercapturic

acid (A'-acetyl-cysteine conjugate) prior to excretion. This is accomplished by renal

y-glutamyltranspeptidase and dipeptidases which cleave the glutathione 5-conjugate to its

corresponding cysteine S'-conjugate. In the kidney, cysteine 5-conjugates are //-acetylated

by A^-acetyl-transferase, forming a mercapturic acid which can be excreted in the urine.

While this type of phase n metabolism serves to detoxify most compounds, several groups

of chemicals may be bioactivated to toxic metabolites by initially forming a glutathione

conjugate (Anders etai, 1988; Vamvakas and Anders, 1990). Several such mechanisms

have been elucidated. Further metabolism of a cysteine S-conjugate by p-lyase, located in

the renal proximal tubules, yields ammonia, pyruvate and reactive thiol compounds which

can form nephrotoxic agents. Certain halogenated alkenes produce renal toxicity by this

mechanism (Koob and Dekant, 1991). Some compounds are also believed to react with

glutathione in a reversible manner. This could allow a conjugate to circulate in a relatively

stable form where it could be reactivated to cause toxicity in other tissues.

Conjugation with glutathione in the liver is a major route of metabolism for certain

PAs (LaFranconi etal., 1985; Lame etal., 1990; Mattocks etai, 1991; Yan and Huxtable,

1994; Lame et ai, 1995). The presence of an AZ-acetyl-cysteine conjugated pyrrole in the

urine of rats treated with monocrotaline or senecionine is evidence of formation of a

cysteine 5-conjugate (Estep et al.. 1990). We have shown that 7-GSDHP formation occurs

Page 109: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

105

in an isolated rat liver preparation perfused with the macrocyclic PAs monocrotaline,

Irichodesmine, seneciphylline and retrorsine (Yan etai., 1995). No evidence of GSH

conjugation was observed when a mixture of the 9-monoesters lycopsamine and

intermedine were perfused in this preparation, as detected by TLC and treatment of bile

with ethanolic silver nitrate. Mattocks et al. {1991) reported a similar occurrence when the

9-monoester heliotrine was tested in the same preparation. This may be due to the low rate

of metabolism of these alkaloids, as well as a lesser reactivity of C9 on the pyrrole nucleus

towards sulfhydryl groups. Formation of 7-GSDHP is likely a result of the non-enzymatic

reaction between the electrophilic DHA and the nucleophilic glutathione. In vitro studies

with hepatic microsomes incubated with senecionine have shown that the addition of GSH

5-transferase (from rat liver cytosol) to the preparation did not increase the rate of 7-

GSDHP formation (Reed et al., 1992). Dehydromonocrotaline reacts rapidly at its 7

position with both cysteine and glutathione (Mattocks et al., 1991). DHP, the hydrolytic

product of DHAs, does not react significantly with glutathione or cysteine at physiological

pH, nor is its ^-conjugation catalyzed by glutathione 5-transferases (Robertson et al.,

1977; Reed et al., 1992). Recent studies have suggested that 7-GSDHP and 7-CYSDHP

may be involved in the extrahepatic toxicity of monocrotaline (LaFranconi et al., 1985;

Huxtable etal., 1990). Although pneumotoxicity is dependent on metabolism, the

metabolite(s) responsible have yet to be identified. Dfff is not responsible for the

pneumotoxicity associated with this PA, and whether the highly reactive

dehydromoncrotaline can survive transport to the lungs is uncertain. This makes die sulfur-

conjugated pyrroles attractive candidates as pneumotoxic metabolites of PAs. They may act

as a transport mechanism, delivering a pyrrolic nucleus to the lungs where it could be

released in a reactive form. In light of recent evidence of glutathione-dependent

bioactivation of xenobiotics, we have investigated the role of 7-GSDHP and the putative

further metabolite, 7-CYSDHP in the toxicity of PAs.

Page 110: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

106

In contrast to earlier studies (Huxtable et al., 1990; Bowers et al., 1991), but in

agreement with anodier (.Pan et al., 1993), 7-GSDHP did not produce toxicities associated

with the PA monocrotaline. Aldiough a small but significant increase in right heart / left

heart ratio was observed, this appears due to the small size of the left heart. Additionally,

because a corresponding increase in lung mass was not observed in these animals, the

increase in right heart size was probably artifactual. The lack of observed toxicity may be

related to the age of the rats studied, as younger rats are clearly more sensitive to PA

intoxication than older ones (Schoental, 1959, 1970; Jago, 1970, 1971; Mattocks and

White, 1973). Huxtable et al. (1990) and Bowers et al. (1991) gave 7-GSDHP to rats

weighing 50-70 g, while in the present study and that by Pan et al. (1993), animals

weighed from 160 to 220 g. The larger rats in this study were used to facilitate the i.v. tail

vein injections.

The putative further metabolite, 7-CYSDHP produced some of the toxicities

associated with monocrotaline. most notably an increase in lung mass. Only a slight

accompanying increase in right heart size was observed, attributable to the relatively small

increase in lung mass. Rats treated with 7-CYSDHP had a small, but significant decrease in

liver weight compared to controls. Because treated groups were smaller than controls,

livers may have been smaller because livers are proportionally smaller in smaller rats

(Mattocks, 1986). Expression of toxicity by 7-CYSDHP required high doses (300 and 600

^imol/kg). Similar changes to the lungs, liver and heart were produced by 7-GSDHP at

these same doses by Bowers etal. (1991) in comparable sized rats. By contrast, greater

pneumotoxicity (as measured by increased lung mass) and right heart hypertrophy is

produced by dehydromonocrotaline at 9-15 ^imol/kg or the parent PA at 185 |j.mol/kg. Yan

and Huxtable (1994) have demonstrated that less than 2% of perfused monocrotaline is

released into the perfusate as 7-GSDHP from isolated rat liver preparation.

Page 111: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

107

Studies indicate that 7-GSDHP is excreted primarily into the bile (LaFranconi et al..

1985; Yan and Huxtable. 1994; Yan ei ai., 1995; Lame ecaL, 1995). Thus the doses of

sulfur-conjugated pyrroles used in this study do not likely reflect their in vivo disposition.

As well, they would be released slowly from the liver rather than as a bolus dose. The

sulfur-conjugated pyrroles are highly water soluble, and do not show alkylating activity.

They are unreactive towards Thiopropyl-Sepharose resin or 4-(/?-nitrobenzyl)pyridine

unless silver nitrate is present to cleave the thioether bond, reforming the pyrrole-stabilized

carbonium ion (Mattocks and Jukes, 1990a). Renal toxicity in rats has been reported with

the PA monocrotaline (Ratnoff and Mirick, 1949; Roth et al., 1981; Kurozumi et al.,

1983). Despite the localization in the kidney of enzymes involved in further metabolism or

generation of toxic metabolites of GSH S'-conjugates, any role of 7-GSDPIP in renal

toxicity is not supported by our findings or those by other researchers. It also does not

appear that 7-GSDHP produces pneumotoxicity at relevant doses. Retrorsine-perfused rat

livers released into the bile four times more, and seneciphylline-perfused livers two times

more 7-GSDHP than monocrotaline-perfiised liver (Yan etal., 1995). Yet the former two

alkaloids are not normally pneumotoxic. Due to the low toxicity and disposition of 7-

GSDHP and 7-CYSDHP, we conclude that hepatic conjugation with glutathione by

metabolites of PAs likely represents a detoxication pathway.

Chemistry of Sulfur-Bound Pyrroles

Thiopropyl-Sepharose 6B resin, a solid phase reagent containing immobilized thiol

groups (anonymous, 1977), is useful in the chromatography of many compounds. The

resin must be "activated" prior to use with a reducing agent such as P-mercaptoethanol.

This uncovers the thiol groups, as they are substituted with a 2-thiopyridyl moiety which

serves as a protective agent. The thiopropyl group is immobilized to the Sepharose matrix

by an ether linkage, and is stable under normal conditions.

Page 112: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

108

It has been established that DHAs react covalently with activated Thiopropyl-

Sepharose resin, forming a stable thioether pyrroie (Mattocks and Juices. 1990b; Mattocks

et ai, 1990). This resin was used to the demonstrate the release of DHAs from isolated

liver perfused with PAs (Mattocks et al., 1990). Glowaz et al. (1992) used the resin to

show production of DHAs by in vitro hepatic microsomal systems incubated with PAs.

Thiopropyl-Sepharose is useful in these studies because of its selectivity for the DHA.

PAs, and metabolites including DHP and 7-GSDHP will not bind to the resin at

physiological pH (Mattocks and Jukes, 1990a, 1990b; Glowaz et al., 1992). As

summarized in Figures 22 and 23, thioether bonds are cleaved in the presence of silver

nitrate or mercuric chloride, solvolizing the pyrrole moiety. Mattocks and Jukes (1990a)

proposed that the metal cation cleaves the sulfide bond, producing a pyrrole stabilized

carbonium ion which reacts with an available nucleophile. This reaction is analogous to the

cleavage of allylic thioethers by silver ion (Saville, 1962). When the pyrrole-bound resin is

treated with silver nitrate in the presence of excess ethanol, a pyrrolic ethyl ether is formed.

This reaction is termed a solvolysis reaction because the solvent, in this case ethanol,

serves as the nucleophile. The position of the ethoxy substitution on the pyrrole represents

the location of the former thioether bond. However, the ethanol must be buffered to prevent

acid-catalyzed alkoxide exchange of hydroxyl substitutions for an ethoxy group (Mattocks

and Jukes, 1990a). Protons released during solvolysis would catalyze diether formation,

preventing true determination of the position of the thioether linkage. A DHA can alkylate

sulfhydryls at either its 7 or 9 position. The subsequent 7- or 9-ethoxy pyrrole is easily

extracted and can be identified by TLC and GC-MS (Table 1 and Figures 23, 39 and 40).

Page 113: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

109

OCOR' ROCO

AgNO

Figure 22 . Proposed reaction of DHAs and Thiopropyl-Sepharose resin. The pyrrole is subsequent solvolized in ethanolic silver nitrate or mercuric chloride, producing an ethoxy pyrrole.

Page 114: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

1 1 0

A) CH

CH

SR

HgCl Nu

OH HO

7,9-Diethoxy DHP

9-Ethoxy DHP 7-Ethoxy DHP

Figure 23. A) Proposed mechanism of cleavage of sulfur-bound pyrroles by mercuric chloride or silver nitrate. From Mattocks (1986). B) Structures of pyrrolic ethyl ethers formed by solvolysis of sulfur-bound pyrroles by ethanolic silver nitrate.

Page 115: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

I l l

This cleavage procedure can be applied to tissue homogenates from animals exposed to

PAs lo establish the position of 5-aikylation in vivo (Mattocks and White, 1970; Mattocks

and Jukes, 1992). Alternatively, by the same procedure, pyrrole bound to resin or tissues

may be released and quantified with Ehrlich Reagent (Mattocks and Jukes, 1990b; Glowaz

et ai, 1992).

Identification of Sulfur-Bound Pyrrole

When resin exposed to macrocyclic DHAs was treated with buffered ethanolic

silver nitrate, 7-ethoxy DHP was produced. 9-Ethoxy or 7,9-diethoxy DHP was not

detected, indicating that the bound pyrrole was in the form of a 7-thioether. In earlier

experiments. Mattocks and Jukes (1990b) presented evidence that when

dehydromonocrotaline was allowed to hydrolyze for 12 sec prior to passing over the resin,

binding occurred at the C9 position, as indicated by the formation of 9-ethoxy DHP from

the resin. Our results do not support the theory that the 9-ester can bind the resin once the

7-ester has hydrolyzed. Further, this is not supported by the results of our in vivo studies,

and studies by others (Mattocks and Jukes, 1990a), as only 7-sulfur adducts were detected

in tissues of animals dosed with macrocyclic PAs. Alkylation at C7 is favored as it forms a

secondary carbonium ion as opposed to the primary carbonium ion formed at C9. The

secondary ion is more stable (the positive charge is delocalized and stabilized by

hyperconjugation with the adjacent alkyl substitutions), and hence forms more easily. The

concept that DHAs can cross-link or bialkylate macromolecules is not supported by these

results, as 7,9-diethoxy DHP has not been detected. Possibly steric factors are involved, as

two large molecular weight nucleophiles would have to be favorably situated to be bound

by the same pyrrole moiety. Only a minimal amount of dehydrolycopsamine bound the

resin, which when treated in the manner above, yielded a trace quantity of 9-ethoxy DHP.

Page 116: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

1 1 2

Wild (1994) observed that the 9-monoester dehydroheliotrine was unreactive towards the

resin.

Relative Reactivity ofDHAs Toward Thiopropyl-Sepharose

We attempted to use the recirculating system coupled with Thiopropyl-Sepharose

resin developed by Mattocks and Jukes (1990b) and described in Methods, for the

determination of aqueous half-lives of DHAs. As mentioned earlier, the system proved

cumbersome and results were difficult to reproduce. Thus the system was superseded by

the use of the potentiometric method. However, some interesting observations were found

with the former system. We examined the relative reactivity of a series of DHAs towards

Thiopropyl-Sepharose resin under aqueous conditions at physiological pH. The reactivity

of DHAs toward the resin is a function of the ability of the DHA to form a carbonium ion

when introduced to the aqueous environment. The quantity of pyrrole binding is dependent

on the competition between water (hydrolysis) and die immobilized sulfhydryls. In half-life

studies, the log quantity of resin bound verse time of hydrolysis (length of delay tubing) is

plotted. The slope represents the rate of hydrolysis. This slope is unaffected by inherent

differences in reactivity towards the resin by DHAs. We found that the system was not

sufficiently sensitive to indicate Oxie differences in 7-ester DHA reactivity, however.

Therefore the studies were terminated in favor of the potentiometric method.

The most important finding was that the 9-monoester dehydrolycopsamine was

only minimally reactive toward the resin. However, the results are consistent with binding

of macrocyclic DHAs to the resin, in which 9-adducts were not detected. Additionally,

sulfur-conjugated pyrroles could not be detected in bile from rat livers perfused with the 9-

monoester PAs heliotrine or lycopsamine (Mattocks etal., 1991; Yan and Huxtable,

unpublished observations). When dehydro-7-acetyl-lycopsamine was tested, it was found

Page 117: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

1 1 3

to be quite reactive towards the resin in comparison to the latter. This demonstrates the

importance of the 7-esler in terms of reactivity towards large molecular weight

nucleophiles.

Hydrolysis of DHAs

The quantity and physicochemical properties of a DHA released into the circulation

from the liver probably determine the nature of extra-hepatic toxicity. Because DHAs are

short-lived, differences in their half-lives may be a factor in the type and extent of observed

toxicity. The aim of these experiments was to establish the aqueous half-lives of a series of

DHAs. The instability of the DHAs makes it impossible to determine their half-lives in

vivo. In vitro determination is also difficult. Mattocks and Jukes (1990b) used the

recirculating system described above to estimate DHA half-lives. The system is

cumbersome, slow, and inaccurate at longer time points. The results are difficult to

reproduce because only a small proportion of the DHAs binds the resin (Table 9). The

potentiometric method we have developed for estimating aqueous half-lives of DHAs is

simple, rapid and reproducible. Ester hydrolysis can be either acid or base catalyzed.

However, the near log-linearity of the results (Figure 12) indicates that the hydrolysis rates

of the DHAs were not significantly affected by the pH of the aqueous medium over the

range used. Thus the use of a pH stat device was unnecessary.

When a DHA enters an aqueous environment, the ester rapidly hydrolyzes by an

1 mechanism to the corresponding pyrrolic alcohol and necic acid. The rate of hydrolysis

is related to the structure of the esterifying acid. Because the departing carboxylate may

remain momentarily attached at the 9 position, there is a possibility the ester may reform,

slowing the rate of hydrolysis. For example, the larger isopropyl group at C14 of

dehydrotrichodesmine 5.36 sec) compared to the methyl of dehydromonocrotaline

(r,/2 = 3.39 sec) could produce greater hindrance to an incoming nucleophile, or could

Page 118: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

1 1 4

cause the departing acid to remain closer to the carbonium ion. This would increase the

probability liial Liie esler would reform, hence slowing hydrolysis. Such a hypothesis

remains consistent with an I mechanism. In support of this observation is the high rate

of hydrolysis of dehydro-7-acetyllycopsamine (^,/2= 0.39 sec), an acyclic diester. Once the

7-acetate dissociates, it is unlikely to reform because it is not attached at the 9 position. The

acyclic 9-monoester lycopsamine 1.02 sec) also hydrolyzes faster than the

macrocyclic DHAs.

The stability of the departing carboxylate may also play a role in DHA reactivity.

The esters of DHAs hydrolyze rapidly because carboxylates are good leaving groups. This

is because in the carboxylate the negative charge of the oxygen anion can be spread over

both oxygen atoms, stabilizing the molecule. In contrast, hydroxyls are poor leaving

groups because they are strongly basic. Their negative charge is localized to one oxygen

and is therefore much less stable. Thus, the weaker the base (or stronger the acid), the

better the leaving group. Synthetic acetyl ester pyrroles are more reactive than

corresponding pivalyl esters (Mattocks, 1986). Acetic acid has a pAT^ of 4.72 compared to

5.01 for pivalic acid. A substituent attached to the carboxyl group which has an electron

withdrawing effect will further stabilize the carboxylate anion (lowering the pATJ, making it

a better leaving group. Dehydroretrorsine (?,/2= 1-06 sec) is significandy more reactive than

dehydroseneciphylline (?,/2= 1-60 sec). Because their structural differences all well

removed from the C7 ester, retronecic acid and seneciphyllic acid may have similar pAT^s.

The reason for their differences in reactivity is unclear. The results of these experiments

clearly establishes the fact the DHAs differ significantly in their rates of hydrolysis. The

relationship of these results and other physicochemical factors to metabolism and toxicity

will be discussed below.

Page 119: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

1 1 5

Alkylation of 4-(/7-nitrobenzy[)pyridine by DHAs

Pyrrolic metabolites of PAs, as well as synthetic pyrroles, have ijeen shown to

alkylate the nucleophile 4-(p-nitrobenzyl)pyridine, forming a colored adduct which can be

detected spectrophotometrically (Figure 24) (Mattocks, 1969; Karchesy and Deinzer, 1981;

Karchesy et al., 1987).

NO-. NO-.

ROCO OCOR

CH- CHi "OH

N^ .OH

NO.

^max 565 nm

Figure 24. Proposed reaction between DHAs and 4-(p-nitrobenzyl)pyridine.

This reaction is useful in the assessment of relative alkylating reactivity of pyrroles

(Mattocks, 1969). Alkylation takes place in aqueous acidified acetone in the presence of

excess nucleophile and is therefore pseudo-first order. DHAs (pyrrolic esters) are more

reactive toward 4-(/7-nitrobenzyl)pyridine than the corresponding alcohol pyrrole, DHP.

Presumably, the pyrrole forms an adduct at its most reactive center by covalently binding

the cyclic nitrogen of 4-(p-nitrobenzyl)pyridine, via an S^l mechanism (Mattocks, 1986).

A second, slower color-forming reaction has been shown to occur as well. Karchesy et al.

(1987) have used mass spectral techniques to show that this reaction (represented by A:,) is

a result of oligomerization reactions by the pyrrole ring, and is not related to the initial

Page 120: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

1 1 6

reactivity of the DHA. Although DHAs are considered bifunctional alkylating agents,

alkylation of the pyrrole at both C7 and C9 is unlikeiy due to the proximity of the positive

charges formed by the 4-(/7-nitrobenzyl)pyridinium ions (Karchesy etal., 1987).

As all die compounds tested were retronecine based (as opposed to heliotridine

based), as with hydrolysis, differences in reactivity of the DHAs with A-{p-

nitrobenzyI)pyridine are a reflection of the different esterifying moieties on die pyrrole

nucleus. Because the same chemical factors which determine hydrolysis rates probably

influence alkylation by 4-(p-nitrobenzyl)pyridine, the results of the two expetiments are

similar. The same rank order of reactivity is observed for the macrocyclic diester PAs

toward 4-(/7-nitrobenzyI)pyridine as toward water (hydrolysis) (Tables 10 and 11).

Karchesy et al. (1987) observed a similar rank order of reactivity for

dehydromonocrotaline, dehydroretrorsine and dehydroseneciphylline as found in diis

study. Dehydro-7-acetyI-lycopsamine reacted at approximately twice the rate of

dehydrolycopsamine in both systems, demonstrating the greater reactivity of the 7-ester

position of the pyrrole compared to the 9-ester. The increased reactivity of C7 compared to

C9 is also demonstrated by 2,3-dihydro-l//-pyrrolizine-l-ol and dehydrosupinidine

(Figure 25). The hydroxyl of the former reacts at 2.5 times the rate with

4-(/7-nitrobenzyl)pyridine as the hydroxyl of the latter (Karchesy et al., 1987).

Figure 25. Structure of 2,3-dihydro-l//-pyrrolizine-l-ol and dehydrosupinidine.

2,3-dihydro-iff-pyiTolizine-l-ol dehydrosupinidine

Page 121: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

1 1 7

This ratio is comparable to dehydro-7-acetyl-lycopsamine and dehydrolycopsamine. While

the reaction between DHAs and 4-(/7-nitrobenzyi)pyridine is not conducted under

physiological conditions, the results provide further support for the view that DHAs differ

significantly in reactivity.

Inhibition of Yeast and Rat Brain Hexokinase by DHAs, in vitro

Hexokinase is a globular protein of approximately 100-kDa, found in microbes,

plants and animals (Wilson. 1995). The enzyme is divided roughly in half by a deep cleft,

with each "subunit" having a mass of approximately 50-kDa. The C-terminus half contains

the binding sites for both ATP and glucose, and is thus considered the catalytic subunit.

Hexokinase catalyzes the first step of the glycolysis pathway, by phosphorylating glucose

to produce glucose-6-phosphate at the expense of ATP. The hexokinases are sulfhydryl-

rich, although the cysteine residues are not well conserved across the family, indicating

they are probably not essential for catalysis (Colowick, 1973; Chou and Wilson, 1974).

Bovine type I hexokinase contains 21 cysteine residues, at least 12 of which exist as free

sulfhydryls. Yeast hexokinases contains at least 8 cysteine residues, but no disulfide

bridges (Wilson, 1995). The yeast enzyme and rat brain enzymes are closely related

(Middleton, 1990). Many studies have shown that reagents reacting with cysteine residues

inactivate hexokinase (Colowick, 1973; Puri etal., 1988; Lenzen etal., 1990; Munday et

al., 1993). The effects of non-specific alkylating agents on yeast hexokinase has also

recently been studied (Puri and Roskoski, 1993; 1994). Because hexokinase is relatively

uncomplicated to assay, and appears sensitive to a variety of alkylating agents, we have

chosen this enzyme as a model to study the interactions between DHAs and an important

cellular macromolecule.

Page 122: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

1 1 8

Yeast and rat brain hexokinase phosphotransferase activity was measured by

following the reduction of NADP to NADPH in a coupied assay with giucose-6-phosphate

dehydrogenase (Figure 26).

Flcxokitiflsc Glucose + ATP :r GIucose-6-Phosphate + ADP

Mg-^

and Glucose-6-Phosphate

Glucose-6-Phosphate + NADP + H.,0 Dehydrogenase ^

6-Phosphogiuconate + NADPH + 2H"^

Figure 26. Schematic of the hexokinase assay. The buildup of NADPH is followed spectrophotometrically at 340 nm.

Yeast hexokinase was rapidly inactivated by 7-ester DHAs, as determined by loss of

phosphotransferase activity of the enzyme. Complete inactivation occurred after a one

minute incubation with 20 mM dehydrotrichodesmine. The rapid time course was expected,

as virtually all alkylating activity of the DHAs would be lost after a one minute exposure to

the incubation solution. That dehydrotrichodesmine, dehydroretrorsine and dehydro-7-

acetyl-lycopsamine had similar IC50 values indicates that alkaloid structure did not effect the

ability of the DHA to inhibit hexokinase. While the actual rate of alkylation may be affected

by DHA structure, the process is too rapid to easily evaluate. The 9-monoester

dehydrolycopsamine was a much weaker inhibitor of hexokinase than

dehydrotrichodesmine. This may be related to the comparatively low alkylating activity of

9-monoesters toward sulfur nucleophiles. as reported in Table 9 and by Wild (1994).

DHP, the hydrolysis product of DHAs, was not active towards hexokinase under the

Page 123: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

1 1 9

studied conditions. DHP is a weak alkylating agent, except under acidic conditions

(Mallocks and Bird, 1983b). The 7.6 pH of the incubation solution would not favor

alkylation by DHP in a short time period. Consistent with their relative unreactivity, parent

PAs had no effect on hexokinase. In parallel assays, the activity of glucose-6-phosphate

dehydrogenase was not inhibited by dehydrotrichodesmine. The Candida utilis (Torula

yeast) form of the enzyme used in these assays is inactivated by various nucleosides, but

not by compounds which modify cysteine residues (Domshke etal., 1970; Levy, 1979).

Why hexokinase, and not glucose-6-phosphate dehydrogenase was inhibited by DHAs is

unknown, but sulfhydryl modification is a possible explanation. At high concentrations,

DHAs are susceptible to rapid polymerization in an aqueous environment, especially under

acidic conditions (Mattocks, 1986). In the 10 and 20 mM DHA incubations with enzyme,

some pyrrole polymerization was visible as a slight pinkish tint to the solution. Because

glucose-6-phosphate dehydrogenase was not inhibited by DHAs, enzyme was not

precipitating with, or "coated" by a pyrrolic polymer (Mattocks, 1986). This suggests that

hexokinase inhibition was not due to such a process, providing support for an alkylation

reaction with DHAs.

Experiments were conducted to generate saturation curves for hexokinase in the

presence and absence of dehydrotrichodesmine (Figure 17). This type of analysis provides

the and for hexokinase under assay conditions. Because dehydrotrichodesmine

caused a decrease in without significantly affecting K^, the inhibition appears to result

from a non-competitive interaction. The effect on the assay system is the same as removing

enzyme; is lowered because it is concentration dependent, while is not affected.

When the data were plotted in double-reciprocal fashion, the regressions were also

suggestive of non-compedtive inhibition (Figure 18). Because the DHAs are powerful

alkylating agents, they likely bind to the enzyme in a irreversible, non-selective manner.

That mannose was only mildly protective against inhibition by dehydrotrichodesmine

Page 124: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

120

suggests that little DHA is reaching the glucose binding site, which is located at the bottom

of a hydrophobic cleft between the two halves (Wilson, 1995). Possibly DHAs are too

large to reach diis site, or do not have time to reach it because of their short half-lives. ATP

provided considerable protection against inactivation by DHAs. This suggests that

alkylation by DHAs at or near the ATP binding site is partially responsible for inhibition.

The ATP binding site has not been conclusively identified, because a crystal structure of

hexokinase-ATP has not been produced. It has been suggested however, that the site is

located on the surface of the C-terminus subunit, near the glucose cleft (Shoham and Stietz,

1980). Such a location could be more accessible to a highly reactive DHA. Alternatively,

ATP may be protecting another important site on the enzyme by inducing a conformational

change, which then does not allow DHAs access to the site.

In fact, these studies do not prove that DHAs are covalently binding to hexokinase.

However, that the parent alkaloid and DHP do not inhibit the enzyme suggests alkylation

by the DHA as a mechanism of inhibition. The many reactive sulfhydryl groups of

hexokinase are potential sites of alkylation (Colowick, 1973; Wilson, 1995). Compounds

that covalently react with these sulfhydryls, such as alloxan (Munday etai, 1993),

inactivate the enzyme. Sulfhydryl modification likely results in an alteration of the

enzyme's conformation or steric obstruction of a catalytic site (Wilson, 1995), as the

cysteine residues are considered non-essential for catalysis (Colowick, 1973). In our

studies, glucose-6-phosphate dehydrogenase was from Candida utilis . This form of the

enzyme contains 8 thiol groups per mole of enzyme, 2 of which are reactive to Ellman's

Reagent without partial denamration (Domshke et al., 1970). The sensitivity of this enzyme

family to sulfhydryl alkylating agents is variable (Levy, 1979). Glucose-6-phosphate

dehydrogenase from C. utilis was shown to be resistant to inhibition by sulfhydryl

alkylating agents such as 5,5'-dithiobis(2-nitrobenzoic acid). The lack of enzyme inhibition

by dehydrotrichodesmine may be related to the low quantity of reactive thiols. Although we

Page 125: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

1 2 1

have not demonstrated that DHAs inactivate hexokinase by sulfhydryl binding, it remains

an attractive hypothesis.

Preliminary experiments show that rat brain hexokinase was inhibited by DHAs in a

similar fashion to the yeast enzyme. All 7-ester DHAs significandy inhibited the enzyme.

The 9-monoester, dehydrolycopsamine, DHP, or parent PA did not inhibit the enzyme. An

interesting finding was that inhibition by dehydrotrichodesmine was increased when the

enzyme was pretreated for 24 hr with P-mercaptoethanol (Figure 21). Sulfhydryl reagents

such as P-mercaptoethanol and cysteine increase the activity of rat brain hexokinase (Figure

21) (Bennet et al., 1962), possibly by reducing disulfides. Whether the increase in

inhibition by dehydrotrichodesmine was associated with additional sulfhydryl binding is

unknown, however. Mild protection was afforded to brain hexokinase by glucose against

inhibition by dehydrotrichodesmine, providing support for the formation of an enzyme-

pyrrole adduct. These experiments demonstrate that DHAs, probably through alkylation,

can adversely affect the activity of a critical cellular component. This does not imply that

inhibition of hexokinase occurs in vivo, yet similar types of interactions might be expected

between DHAs and macromolecules.

The Relationship Between the Physicochemical Properties of DHAs and

Their Further Metabolism and Toxicity

Macrocyclic diester PAs

Results from this dissertation, as well as odier studies in our laboratory, have

identified physicochemical properties associated with the quantitative and qualitative

differences in toxicity of PAs (Yan and Huxtable, 1995; Yan etai, 1995; Huxtable etal.,

1996). The toxicity produced by PAs is a consequence of their hepatic metabolism to

DHAs, which alkylate cellular nucleophiles. Lethality and specific organs affected are

Page 126: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

122

related to both the rate of formation of DHAs, and the pattern of their further metabolism.

As discussed earlier, rate of pyrrole foniiation is primaiily a function of PA lipophilicily

and an acid structure which favors dehydrogenation rather than hydrolysis and A/-oxidation

(Mattocks, 1986). From studies in this laboratory (Yan et al., 1995), Table 14 shows

differences in rates of pyrrole formation in the isolated rat liver perfused with PAs.

Consistent with other studies, retrorsine was the most extensively metabolized macrocyclic

PA (Mattocks, 1972, 1986). Of the two primarily hepatotoxic PAs, significantly more

bound pyrrole was detected in the liver following perfusion with retrorsine than

seneciphylline. The higher rate of pyrrole formation may account for the lower LDjg and

greater hepatotoxicity (Culvenor et al., 1975) of retrorsine compared to seneciphylline.

Mattocks (1972) has demonstrated a strong correlation between hepatotoxicity and the

quantity of liver-bound pyrrole following PA exposure. Wild (1994) has shown that when

PAs are administered to rats by i.p. injection, significantly more liver-bound pyrrole is

found following retrorsine injection than either monocrotaline or trichodesmine. This is in

keeping with the greater hepatotoxicity and higher rate of metabolism of retrorsine than

monocrotaline (Culvenor et al., 1975). While the hepatotoxicity of trichodesmine has not

been assessed, its metabolism is associated with significantly less liver-bound pyrrole than

monocrotaline (Table 15), despite the faster metabolism of the former (Yan et al., 1995;

Huxtable er a/., 1996).

In addition to rate of pyrrole formation, toxicity of PAs is related to the pattern of

further metabolism of the DHA. This pattern is in turn related to DHA stability, particularly

at the 7-ester position. As shown in studies with the isolated perfused liver (Table 14), the

more resistant a DEIA is to nucleophilic attack (the longer the half-life), the greater the

proportion that is released from the liver, increasing the availability of DHA for extrahepatic

reactions (Yan et al., 1995). The more reactive DHAs of retrorsine and seneciphylline favor

hydrolysis or nucleophilic attack closer to their sites of formation. Thus, a greater

Page 127: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

123

proportion of the total pyrrole produced is released from the isolated liver as non-toxic 7-

GSDHP, and a smaller proportion of DHA is released into the perfusate compared to

monocrotaline and trichodesmine. Quantitatively, the amounts of DHA released into the

perfusate by retrorsine, seneciphylline and monocrotaline perfused livers were all similar

(but represent a variable proportion of the total pyrrole produced). Monocrotaline is the

only one of these PAs associated with extrahepatic toxicity, suggesting that of the

corresponding DHAs, only dehydromonocrotaline is sufficiently stable to reach the lungs

in quantities necessary to produce toxicity. Wild (1994) demonstrated that monocrotaline-

treated rats had more lung pyrrole than retrorsine-treated animals. Following exposure to

the PAs, four times as much DHA is released from the liver perfused with trichodesmine

than monocrotaline. This is related to the higher rate of metabolism of trichodesmine and

the longer half-life of its DHA (Table 14). Consequently, in trichodesmine-treated animals,

more DHA will be available to produce toxicity away from the liver. Such a conclusion is

supported by the observed pattern of tissue alkylation in vivo 18-24 hr after i.p. injection of

the PAs. Following a dose of PA approximately 60% of the LDjq, trichodesmine-treated

rats had significantly higher brain pyrrole than monocrotaline-treated animals (Table 15)

(Huxtable et al., 1996). When tested on an equimolar basis, brain pyrrole was ten-fold

higher in the trichodesmine-treated animals (Wild, 1994). The differences in brain pyrrole

levels are probably a reflection of greater delivery and lipophilic character of

dehydrotrichodesmine. Due to the isopropyl substimtion at CI4, trichodesmine has a

greater lipid:aqueous partition coefficient than monocrotaline (Huxtable etaL, 1996).

Pyrrole was not detected in brains of retrorsine-treated rats (Wild, 1994), in keeping with

the very short half-life of its DHA. Consistent with its pneumotoxicity, monocrotaline-

treated animals had higher lung pyrrole levels than those administered trichodesmine (Table

15). The slower reactivity of dehydrotrichodesmine and dehydromonocrotaline with water

or GSH ensures a greater proportion of DHA will be released from the liver into the

Page 128: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

124

circulation (Table 14). This stability also allows DHA to reach sites downstream from the

liver to produce extraliepatic toxicity. Tlie higher rate of tnelaboHsm of Irichodesmine

compared to monocrotaline is attributable to the isopropyl grouping of the former. The

increased lipophilicity may make it a better substrate for P450, or the bulkier acid may be

more resistant to hydrolysis by hepatic esterases.

Open ester DHAs

The low toxicity of PA monoesters such as lycopsamine (LD50 > 3 mmol/kg

(Fowler et al., 1967)) can also be related to its structural properties. Monoesters are not

metabolized to pyrroles by hepatic P450s (Glowaz et al. 1992, Yan et al., unpublished

observations) as extensively as diester PAs, which may be related to their hydrophilicity.

thus promoting excretion. Retronecine based monoesters, including lycopsamine, are also

susceptible to hydrolysis by esterases. Once metabolized to a pyrrole, open ester DHAs

appear to have very short half-lives (Cooper and Huxtable, 1996), consistent with the

observation that they only produce hepatotoxicity (Mattocks, 1986). Additionally, our

studies indicate that 9-monoester DHAs are poor alkylating agents of sulfur

macromolecules under physiological conditions, although the reason for this selectivity is

unknown.

The hepatotoxicity and lethality of open diester PAs are highly variable (Mattocks,

1986). Because they all contain a 7-ester, their corresponding DHAs are strong alkylating

agents. The relative toxicity of the diesters is primarily dependent on the structure of the

esterifying acid at CI. Larger, branched acids such as tiglic or angelic are more resistant to

hydrolysis by esterases than smaller acids such as acetic (Mattocks, 1982). Culvenor et al.

(1975) found 7-acetyl-heliotrine to be equally hepatotoxic as heliotrine, which he attributed

to loss of the 7-acetyl group to hydrolysis. The more resistance to hydrolysis, the greater

Page 129: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

125

Table 14

Relationship Between Pattern of Metabolism of PAs in the Isolated Perfused Rat Liver and DHA Half-Life"

PA % Metabolized 7-GSDHP^ DHA''

(with LDso) to Pyrrole* (% of total pyrrole) (% of total pyrrole) r,^of DHA

Retrorsine (97 nmol/kg)

31.3 54.4 ± 2.7 10.2 ±0.43 1.06 ±0.20

Seneciphylline (230 |imol/kg)

17.0 45.3 ±3.7 18.5 ±2.1 1.60 ±0.28

Monocrotaline (335 |imol/kg)

9.7 36.4 ±3.2 23.4 ± 4.0 3.39 ± 0.99

Trichodesmine (57 |imol/kg)

15.8 9.6 ± 0.8 56.2 ± 7.7 5.36 ± 0.92

Terfusion data taken from Yan etal. (1995). Isolated rat livers were perfused with PAs at 0.5 mM for I hr. Pyrrolic metabolites that were released into the perftisate and bile, or bound to the liver, were quantified with Ehrlich Reagent.

^Percentage of total perfused PA metabolized to pyrrole over the one hour period. '7-GSDHP released into the bile, expressed as a percentage of total pyrrolic metabolites detected.

"IDHA released into the perfusate, expressed as a percentage of the total pyrrolic metabolites detected.

Page 130: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

126

Table 15

Tissue-Bound Pyrrolic Metabolites 24 hr After i.p. Injection of PAs in Rats

Parent Dose Tissue bound pyrrole (nmol/g tissue)

Alkaloid" (jimol/kg) Liver Lung Brain

Monocrotaline 65 17.1 ±0.9 10.4 ± 0.9 1.74 ±0.6

Trichodesmine 15 6.7 ± 1.4 7.7 ± 0.4 3.74 ± 0.3

"Doses represent approximately 60% of the acute LDj^ (Mattocks, 1986; Wild, 1994). Values for each tissue are significantly different at p < 0.05 (From Huxtable nt al., 1996).

Page 131: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

127

the quantity of PA available for metabolism to pyrrole. PAs such as lasiocarpine and

symphytine, have larger, branched esterifying acids at C7 and are quite lipophilic (Mattocks

and Bird, 1983a; Cooper et al., 1996). Because of extensive metabolism to a DHA with a

reactive 7-ester, these PAs are as lethal as some macrocyclics, with LD50S of 340 and 185

|j.mol/kg for symphytine and lasiocarpine, respectively (Bull et ciL, 1958; Hirono et ai.

1979).

Conclusion

This research has investigated how certain physicochemical properties of PA

metabolites relate to the toxicity produced by the parent alkaloids. The relationship between

structure and toxicity is complex and only partially understood. The results of this

dissertation suggest that differences in stability of macrocyclic DHAs are associated with

differences in the pattern of metabolite release observed in the isolated, perfused liver. In

turn, the pattern of metabolite release is related to the quantitative (lethality) and qualitative

(organs affected) toxicity.

Using a potentiometric method, we have shown that the Senecio type DHAs,

dehydroseneciphylline and dehydroretrorsine, are more reactive in an aqueous environment

than the Crotalaria alkaloids, dehydromonocrotaline and dehydrotrichodesmine. Reactivity

of these DHAs involves nucleophilic substitution at the 7 position of the ring system by an

S^I mechanism. As demonstrated in the isolated, perfused liver, the more reactive the

DHA, the greater the proportion reacting with GSH, the greater the proportion alkylating

liver macromolecules, and the lower the proportion released as DHA into the circulation.

Lower quantities of DHAs (with short half-lives) escaping the liver decreases the

probability that other organs will be affected. On the other hand, the more stable DHAs of

monocrotaline and trichodesmine are associated with lower proportions of 7-GSDHP

release and liver alkylation, and higher proportions of DHA released into the circulation.

Page 132: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

128

The higher proportion of the latter increases the risk of extrahepatic toxicity, as greater

quantities of DHA will be delivered to extrahepatic sites. The high iiposoiubiiity, longer

half-life, and greater rate of formation of dehydrotrichodesmine compared to

dehydromonocrotaline favor increased access of DHA to the brain for

dehydrotrichodesmine.

Other contributions of this research include the development of an efficient method

of separating and purifying PAs using high-speed counter-current chromatography. This

technique is useful in supplying the large quantities of PAs necessary for chemical and

toxicological studies. We have also shown that glutathione and cysteine-conjugated

pyrroles have low toxicity, and probably represent a major detoxication pathway for PAs.

This provides further support for DHAs as the primary toxic metabolite of PAs. Overall,

the results of this research gives more insight into some of the toxic properties of certain

types of PAs. In general, this work provides further information on the relationship

between hepatic metabolism of xenobiotics and extrahepatic toxicity.

Page 133: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

129

APPENDIX A

Electron Impact-Mass Spectra of PAs

43

53

120

93

136

157 164

\ / 194

236 I

256 281 325 (M-")

40 80 120 160 200

m/z

240 280 320 360

Figure 27. Electron impact-mass spectrum of monocrotaline.

Page 134: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

130

100

90

80 -3

70

60

50

40

30

20

10

0 J

222 206 /

14.

264

J. 338

\

353 (M^-)

—I—1 r-—f—I—r-—I 1—r— 1

40 80 120 160 200 240 280 320 360

m/z

Figure 28. Electron impact-mass spectrum of trichodesmine.

Page 135: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

1 3 1

100

90 H

80 -

70 A g ^ 6 0 - 3 •w 'cfl C ''J sr, tS 50 -

•s 40 A

13 0^ 30

20

10

0 J

206 250

-Lull

292

\ 337 (M"'

I

T X

160 200 m/z

240 280 320 360

Figure 29. Electron impact-mass spectrum of incanine.

Page 136: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

132

220

246

—r"—'—I— 240 280

320 351 (M-")

—I 320 160 200

m/z

360

Figure 30. Electron impact mass-spectrum of retrorsine.

Page 137: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

133

100 q 136

90

80

^ 70

§ 60 c (U

•S 50 I

"S Oj 40 ^

30

20

10

0

\20

1151

246

jiL

218

L

288 289 333 (M^)

\ , / /

I ' ' ' '"I "" ' I I I

40 80 120 160 200 240 280 320 360

m/z

Figure 31. Electron impact mass-spectrum of seneciphylline.

Page 138: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

100 120

93

80 -

.-s 60 c u

ta 40 -

20

0 J

53

I—^

40

80

80

106

136

151 220 289 291 335 (M^)

1' \ I \ i i

120 160 200 240 280 320 360

m/z

Figure 32. Electron impact-mass spectrum of senecionine.

Page 139: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

135

100 n

80

^ 60

c D

^ 40 ga (U

20 •

69

55 110

122

94

150

168

184

250 304 336

238

264

281

40 80 120

r" ̂ 1>—^-"1——

160 200 240 280 320

423 (m-^) 364 \ /

Jlj—, I I .

360 400 440

m/z

Figure 33. Electron impact mass-spectrum of florosenine.

Page 140: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

138

90 -i

80 H

43

93

80

67

lU

120

LjkJlL

156

40 80 1 1—

120 160 200

254 299(m

240

—I ̂

280

m/z

Figure 34. Electron impact-mass spectrum of lycopsamine/intermedine.

Page 141: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

137

100 1 138

43

93

67 120

i ii In. j.ii jij 1 iwtL 156 181

Jj •I 198 J H-

160 200

m/z

255 298 —l-

341 (m"")

T-

40 80 120 240 280 320 360

Figure 35. Electron impact-mass spectrum of 3-acetyl-lycopsamine/-intermedine-

Page 142: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

100 T

90

80 -

g 70 -

'33 60 -c <u •w c

•Z 5 0 ^

i 40 (U " 0^

30 .

20 r

10 r

0 -

180

43

i| li jflllj

120

93

80 136

L

198

+ 281 296 34 j

\ I \

40 80 120 160 200 240 280 I

320 360

m/z

Figure 36. Electron impact-mass spectmm of 7-acetyl-lycopsamine/-intermedine.

Page 143: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

139

100 T 180

90

80 -

70

60 H

C/D S 50

u 40

-4-* — <u 0^ 30

20

10

0

93

'T-80

120

136

198 381 (m-^)

291 341 J L-

120 160 200 240 280 320 360 400

m/z

Figure 37. Electron impact-mass spectrum of 3,7-diacetyl-lycopsamine/-intermedine.

Page 144: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

100 T 220

90 -

80

70

60 -

C/3

I 50

•>: 40 -

30 -

20 -

10 H

0 J

43 55

93

I

83

67

120

40 80

136 I

141

/

120 160

m/z

238

200 240

281

280

Figure 38. Electron impact-mass spectrum of symphytine and isomers (symlandine, myoscorpine and echiumine). Because a molecular ion was not observed, mass was confirmed by flow-injection APCI.

Page 145: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

100 T 136

90 T

80 -

70 H

35 60 ••

50 -HJ >

13 40 -

30

20 -

10 -

0

29 41

80

51 68

III., Jill

40 80

106

1 1 8

120

152 I

181 (M^-)

160 200

m/z

Figure 39. Electron impact-mass spectrum of 7-ethoxy DHP.

Page 146: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

142

100

90

80 -

70 -

'ES 60 -i

^ 50 -

S 40 -

30 •

20 -

10

0

29

53

39

94

79

65

118

106

136

163

152

181 (M-")

20 40 60 80 100 120 140 160 180 200

m/z

Figure 40. Electron impact-mass spectrum of 9-ethoxy DHP.

Page 147: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

100 n

90

80 ^

70

60

50

40

30

20

10

0 J

120

29

53 93 106

79

1 ii. • .111 (1 i

134

40 80 120

m/z

164

160

209 (M-^)

JU L

200

Figure 41. Electron impact-mass spectrum of 7,9-diethoxy DHP.

Page 148: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

100

90 -

80

70 •

£ ^ 60 • c/a C o a

<u

•s 40 -— 13

30 -

20 -

10

0 J r

154 matrix

136 matrix

107

120

120

x l O — ^

219

207

171 189

n—

160

•T——r-'

200

239

257 (M+H)-"

307 273 matrix

289 matrix

326

240 280

-I—' 1—

320

m/z

Figure 42. Positive ion FAB mass spectrum of 7-CYSDHP. Matrix was 3-NBA.

Page 149: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

145

REFERENCES

Adams, R. aiid Rogers, E.F. (1939) The structure of monocrolaline, the alkaloid in Crotalaria spectabilis and Crotalaria retusa. J. Am. Chem. Soc. 61, 2815-2819.

Adams, R. and Govindachari, T.R. (1949) Senecio alkaloids: The alkaloids of Senecio douglasii, carthamoides, eremophiliis, ampullaceus and parksii. J. Am. Chem. Soc. 71, 1956-1960.

Adams, R. and Gianturco, M. (1956) Senecio alkaloids: the structure of trichodesmine. J. Am. Chem. Soc.1%, 1922-1925.

Anders, M.W., Lash, L.H., Dekant, W., Elfarra, A.A. and Dohn, D.R. (1988) Biosynthesis and biotransformation of glutathione 5-conjugates to toxic metabolites. CRC Crit. Rev. Toxicol. 18, 311-341.

Anonymous (1977) Thiopropyl-Sepharose® 6B. Pharmacia Technical Bulletin Vastros Aros Trycken AB, Sweden.

Asibal, C.F., Gelbaum, L.T. and Zalkow, L.H. (1989) Pyrrolizidine alkaloids from Heliotropium rotundifolium. J. Nat. Prod. 52, 726-731.

Asibal, C.F., Glinski, J.A., Gelbaum, L.T. and Zalkow, L.H. (1989) Pyrrolizidine alkaloids from Cynoglossiim creticiim. Synthesis of the pyrrolizidine alkaloids echinatine, rinderine, and analogues. J. Nat. Prod. 52, 109-118.

Black, D.N. and Jago, M.V. (1970) Interaction of dehydroheliotridine, a metabolite of heliotridine based pyrrolizidine alkaloids, with native and heat-denatured deoxyribonucleic acid in vitro. Biochem. J. 118, 347-353.

Bennet, E.L., Drori, J.B., Krech, D., Rosenweig, M.R. and Abraham, S. (1962) Hexokinase activity in brain. J. Biol. Chem. 237, 1758-1763.

Boppre, M. (1986) Insects pharmacophagously utilizing defensive plant chemicals (pyrrolizidine alkaloids) Naturwissenschaften 73, 17-26.

Bras, G., Jelliffe D.B. and Stuart, K.L (1954) Veno-occlusive disease of liver with nonportal type of cirrhosis, occuring in Jamaica. Arch. Pathol. 57, 285-300.

Bras, G. and Hill, K.R. (1956) Veno-occlusive disease of the liver - essential pathology. Lancet!, 161-163.

Bras, G., Berry, D.M. and Gyorgi, P. (1957) Plants as etiological factors in veno-occlusive disease of the liver. Lancet 1, 960-962.

Bruner, L.H., Hilliker, K.S. and Roth, R.A. (1983) Pulmonary hypertension and ECG changes from monocrotaline pyrrole in the rat. Am. J. Physiol. 257, H300-H306.

Bull, L.B., Dick. A.T. and McKenzie, J.S. (1958) The acute toxic effects of heliotrine and lasiocarpine, and their A^-oxides, on the rat. J. Pathol. Bacteriol. 75, 17-25.

Page 150: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

146

Bull, L.B., Culvenor, C.C.J, and Dick, A.T. (1968) The Pyrrolizidine Alkaloids North-Holland Publ., Amsterdam.

Butler, W.H., Mattocks, A.R. Barens, J.M. (1970) Lesions in the liver and lungs of rats given pyrrole derivates of the pyrrolizidine alkaloids. J. Pathol. 100, 169-175.

Candrian, U., Luthy, J. and Schlatter, C. (1985) In vivo covalent binding of retronecine-labelled [^H] seneciphylline and [^H] senecionine to DNA of rat liver, lung and kidney. Chem.-Biol. Interact. 54, 57-69.

Carballo, M., Mudry, M.D., Larripa, I.B., Villamil, E. and D'Aquino, M. (1992) Genotoxic action of an extract of Heliotropium curassaviciim var. argentinuni. Mutation Res. 279, 245-253.

Cava, M.P., Rao, K.V., Weisbach, J.A., Raffauf, R.F. and Douglas, B. (1968) The alkaloids of Cacalia floridana. J. Org. Chem. 33, 3570-3573.

Chizzola, R. (1994) Rapid sample preparation technique for the determination of pyrrolizidine alkaloids in plant extracts. J. Chromatogr. A 668,427-433.

Chou, A.C. and Wilson, J.E. (1974) A study of the sulfhydryl groups of rat brain hexokinase. Arch. Biochem. Biophys. 163. 191-199.

Chung, W. and Buhler, D.R. (1995) Major factors for the susceptibilty of guinea pig to the pyrrolizidine alkaloid jacobine. Drug Metab. Dispos. 23, 1263-1267.

Chung, W., Miranda, C.L. and Buhler, D.R. (1995) A cytochrome P4502B form is the major bioactivation enzyme for the pyrrolizidine alkaloid senecionine in guinea pig. Xenobiotica 25, 929-939.

Colowick, S.P. (1973) The Hexokinases. In: The Enzymes, vol. 9. Boyer, P.D. (Ed.). Academic, New York. pp. 1-48.

Conway, W.D. (1990) Counter Current Chromatography: Apparatus, Theory and Applications. VCH, New York.

Cooper, R.A., Bowers, R.J. and Huxtable, R.J. In vivo toxicity of thioether metabolites of the pyrrolizidine alkaloid, monocrotaline. (Abstract) Toxicon 31: 121, 1993.

Cooper, R.A., Bowers, R.J., Beckham, C.J. and Huxtable, R.J. (1996) Preparative separation of pyrrolizidine alkaloids by high-speed counter current chromatography. J. Chromatogr. A., in press.

Cooper, R.A. and Huxtable, R.J. (1996) A simple procedure for determining the aqueous half-lives of pyrrolic metabolites of pyrrolizidine alkaloids. Toxicon, in press.

Culvenor, C.C.J, and Smith, L.W. (1966) The alkaloids of Amsinckia species: A. intermedia Fisch. & Mey., A. hispida (Ruiz. & Pav.) Johnst. and A. lycopsoides Lehm. Aust. J. Chem. 19, 1955-1964.

Page 151: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

147

Culvenor, C.C.J., Edgar, J.A., Smith L.W. and Tweeddale, H.J. (1970) Dihyropyrrolizines HI. Preparation and reactions of derivatives related to pyrrolizidine alkaloids. Aust. J. Chern. 23, 1853-1867.

Culvenor, C.C.J., Edgar, J.A., Frahn, J.L. and Smith, L.W. (1980) The akaloids of Symphytum uiplandicum (Russian Comfrey). Aust. J. Chem. 33, 1105-1113.

Culvenor, C.C.J., Edgar, J.A. and Smidi, L.W. (1981) Pyrrolizidine alkaloids in honey from Echium plantagineum L. 29, 958-960.

Culvenor, C.C.J., Clarke, M.,. Edgar, J.A ., Frahn, J.L., Jago, M.V., Peterson J.E. and Smith, L.W. (1980) Structure and toxicity of the alkaloids of Russian Comfrey {Symphtyum Kuplandicum Nyman), a medicinal herb and item of human diet. Experientia 36, 377-379.

Culvenor, C.C.J., Edgar, J.A. and Smith L.W. (1981) Pyrrolizidine alkaloids in honey from Echium plantagineum L. J. Agric. Food Chem. 29, 958-960.

Deinzer, M.L., Thomson, P.A., Burgett, D.M. and Isaacson, D.L. (1977) Pyrrolizidine alkaloids: their occurance in honey from tansy ragwort {S. jacobaea). Science 195,497-499.

Dickinson, J.O., Cooke, M.P., King, R.R. and Mohamed, P.A. (1976) Milk transfer of pyrrolizidine alkaloids in cattle. J. Am. Vet. Med. Assoc. 169, 1192-1196.

Dickinson, J.O. (1980) Release of pyrrolizidine alkaloids into milk. Proc. West. Pharmacol. Soc. 23, 377-379.

Domschke, W., Hinuber, C.V. and Domagk, G.F. (1970) Studies on protein structure and enzyme activity Vni; cysteinyl residues in glucose-6-phophate dehydrogenase from Candida utilis. Hoppe-Seyler's Z. Physiol. C/zem. 351, 194-196.

Driver, H.E. and Mattocks, A.R. (1984) The toxic effects in rats of some sythanecine carbamate and phosphate esters analogous to hepatotoxic pyrrolizidine alkaloids. Chem.-Biol. Interact. 51, 201-218.

Dueker, S.R., Lame, M.W. and Segall, H.J. (1992) Hydrolysis of Pyrrolizidine Alkaloids by Guinea Pig Hepatic Carboxylesterases. Toxicol. Appl. Pharmacol. 117, 116-121.

Dueker, S.R., Lame, M.W., Jones, A.D., Morin, D., Segall, H.J. (1994) Glutathione conjugation with the pyrrolizidine alkaloid, jacobine. Biochem. Biophys. Res. Comm. 198, 516-522.

Dueker, S.R., Lame, M.W. and H.J. Segall (1995) Hydrolysis rates of pyrrolizidine alkaloids derived from Senecio jacobeae. Arch. Toxicol. 69, 725-728.

Eastman, D.F., Dimenna, G.P. and Segall, H.J. (1982) Covalent binding of two pyrrolizidine alkaloids, senecionine, and seneciphylline to hepatic macromolecules and their distribution, excretion and transfer into milk of lactating mice. Drug Metab. Disp. 10, 236-240.

Page 152: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

148

Estep, J.E., Lame, M.W., Jones, D. and Segall, H.J. (1990) A^-acetylcysteine-conjugated pyrrole identified in rat urine following administration of two pyrrolizidine alkaloids, monocrotaiine and senecionine. Toxicol. Lett. 54, 61-69.

Fischer, N., Weinreich B., Nitz, S. and Drawert, F. (1991) Applications of high-speed counter-current chromatography for the separation and isolation of natural products. J. Chromatogr. 538, 193-202.

Frei, H., Liithy, J., Brauchli, J., Zweiffel, U., Wurgler, F.E. and Schlatter, C. 0992) Structure/activity relationships of the genotoxic potencies of sixteen pyrrolizidine alkaloids assayed for the induction of somatic mutation and recombination in wing cells of Drosophila melanogaster. Chem.-Biol. Interact. 83, 1-22.

Fowler, M.E. and Schoental, R. (1967) Toxicity of pyrrolizidine alkaloids fvom Amsinckia intermedia. J. Am. Vet. Med. Assoc. 150, 1305.

Fox, D.W., Hart, M.C., Bergeson, P.S., Jarrett, P.B., Stillman, A.E. and Huxtable, R.J. (1978) Pyrrolizidine (Senecio) intoxication mimicking Reye's syndrome. /. Pediatr. 93. 980-982

Glowaz, S.L., Michnika, M. and Huxtable, R.J. (1992) Detection of a reactive pyrrole in the hepatic metabolism of the pyrrolizidine alkaloid, monocrotaiine. Toxicol. Appl. Pharmacol. 115, 168-173.

Gillis, C.N., Huxtable, R.J. and Roth, R.A. (1978) Effect of monocrotaiine pretreatment of rats on removal of 5-hydroxytryptamine and norepinephrine by perfused lung. Br. J. Pharmacol. 63, 435-443.

Hill, K.R. and Rhodes, K., Stafford, J.L. and Aub, R. (1953) Serous hepatosis: a pathogenesis of hepatic fibrosis in Jamaican children. Br. Med. J. 2, 117-122

Hincks, J.R., Kim, H., Segall, H.J., Molyneux, R.J., Stermitz, F.R., Coulombe, Jr., R.A. (1991) DNA cross-linking in mammalian cells by pyrrolizidine alkaloids: structure-activity relationships. Toxicol. Appl. Pharmacol. Ill, 90-98.

Hirono, I., Mori, H. and Haga, M. (1978) Carcinogenic activity of Symphytum officinale. J. Natl. Cancer Inst. 61, 865-869.

Hirono, I., Haga, M., Fujii, M., Matsuura, S., Matsubara, N., Nakayama, M., Furaya, T., Hikichi, M.. Takanashi, H., Uchida, E., Hosaka, S. and Ueno, I. (1979) Induction of hepatic tumors inrats by senkirkine and symphytine. J. Natl. Cancer Inst. 63, 469-471.

Hsu, I.e., Allen, J.R. and Chesney, C.F. (1973) Identification and toxicological effects of dehydroretronecine. a metabolite of monocrotaiine. Proc. Soc. Exp. Biol. Med. 144, 834-838.

Huxtable, R.J., Ciaramitaro, D. and Einstein, D. (1978) The effect of a pyrrolizidine alkaloid, monocrotaiine, and a pyrrole, dehydroretronecine, on the biochemical functions of the pulmonary endothelium. Molec. Pharmacol. 14, 1189-1203.

Page 153: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

149

Huxtable, R.J. (1979) Pyrrolizidine alkaloids and the lung endothelium: a paradigm of lung damage resulting from circulating toxins. Pyrrolizidine fSenecioj alkaloids: toxicity, metabolism and poisonous plant control measures. Cheeke, P.R., ed. Oregon State University, Corvalis. p. 43-56.

Huxtable, R.J., Luthy, J. and Zweiffel, V. (1986) Toxicity of comfrey-pepsin preparations. New Engl. J. Med. 315, 1095.

Huxtable, R.J. (1989) Human health implications of pyrrolizidine alkaloids and herbs containing them. In: Toxicants of Plant Origin, Vol I: Alkaloids, edited by P.R. Cheeke. Boca Raton, Florida: CRC Press, p 41-86.

Huxtable, R.J. (1990a) Activation and pulmonary toxicity of pyrrolizidine alkaloids. Pharmac. Ther. 47, 371-389.

Huxtable, R.J. (1990b) The harmful potential of herbal and other plant products. Drug Safety 5 (Suppl. I), 126-136.

Huxtable, R.J., Bowers, R., Mattocks, A.R. and Michnika, M. (1990) Sulfur conjugates as putative pneumotoxic metabolites of the pyrrolizidine alkaloid, monocrotaline. Biological Reactive Intermediates IV. Witmer, C.M. et al., eds. Plenum Press, New York, pp. 605-612.

Huxtable, R.J., Yan, C.C., Wild, S., Maxwell, S. and Cooper, R. (1996) Physicochemical and metabolic basis for the differing neurotoxicity of the pyrrolizidine alkaloids, trichodesmine and monocrotaline. Neurochem. Res. 21, 141-146.

Ismailov, N.I. (1970) Heliotropic Toxicosis (Toxic Hepatitis with Ascites). Academy of Sciences of Uzbekistan. Tashkent.

Ismailov, N.I., Madzhidov, N.M., Magrupov, A.L., Makhkamov, G.M. and Mukminova, S.G. (1970) Klinika, diagnostika I lecheniye, trichodesmotokikoza (alimentamo toksichekogo entsefalita) [Clinical signs, diagnosis and treatment of Trichodesma toxicosis (alimentorv toxic encephalopathy)]. Meditsina. Tashkent, Uzbek S.S.R. p. 85.

Ito, Y., Weinstein, M.A., Aoki, I., Harada, R., Kimura, E. and Nunogaki, K. (1966) The coil planet centrifuge. Nature (London) 212, 985-987.

Ito, Y., Sandlin, J. and Bowers, W.G (1982) High-speed prepartive counter-current chromatography with a coil planet centrifuge. J. Chromatogr. 244, 247-258.

Jago, M.V. (1969) The developement of hepatic megalocytosis of chronic pyrrolizidine alkaloid poisoning. Am. J. Pathol. 56, 405-421.

Jago, M.V. (1970) A method for the assesment of the chronic hepatotoxicity of pyrrolizidine alkaloids. Aust. J. Exp. Med. Sci. 48, 93-103.

Jago, M.V. (1971) Factors effecting the chronic hepatotoxicity of pyrrolizidine alkaloids. J. Pathol. 105. 1 - 1 1 .

Page 154: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

150

Jago, M.V., Edgar, J.A., Smith, L.W. and Culvenor, C.C.J. (1970) Metabolic conversion of heliotridine-based pyrrolizidine alkaloids to dehydroheliotridine. Mol. Pharmacol. 6, 402-406.

Johnson, A.E., Molyneux, R.J. and Stuart, L.D. (1985) Toxicity of Riddell's groundsel (Senecio riddellii) to cattle. Am. J. Vet. Res. 46, 577-582.

Kay, J.M. and Heath, D. (1969) Crotalaria spectabilis: The Pulmonary Hypertension Plant. Thomas: Springfield, IL.

Karchesy, J.J. and Deinzer. M.L. (1981) Kinetics of alkylation reactions of pyrrolizidine alkaloid derivatives. Heterocycles 16, 631-635.

Karchesy, J.J., Arbogast, B. and Deinzer, M.L. (1987) Kinetics of alkylation of pyrrolizidine alkaloid pyrroles. J. Org. Chem. 52, 3867-3872.

Kedzierski, B. and Buhler, D.R. (1986) The formation of 6,7-dihydro-7-hydroxy-l-hydroxymethyi-5/f-pyrrolizine, a metabolite of pyrrolizidine alkaloids. Chem.-Biol. Interact. 57, 217-222.

Kelley, R.B. and Seiber. J.N. (1992) Pyrrolizidine alkaloid chemosystematics in Amsinckia. Phytochemistry 31. 2369-2387.

Kearney. T.H. and Peebles, R.H. (I960) Arizona Flora; Second Edition. University of California Press, Berkeley, California.

Kim, H-K., Stermitz. F.R. and Coulombe, R.A. (1995) Pyrrolizidine alkaloid-induced DNA-protein cross-links. Carcinogenesis 16, 2691-2697.

Koob, M. and Dekant, W. (1991) Bioactivation of xenobiotics by formation of toxic glutathione conjugates. Chem.-Biol. Interact. 77, 107-136.

Kurozumi, T., Tanaka, K., Kido, M and Shoyama, Y. (1983) Monocrotaline induced renal lesions. Exp. Mol. Pathol. 39, 377-389.

LaFranconi, W.M. and Huxtable, R.J. (1983) Changes in angiotensin converting enzyme in lungs damaged by the pyrrolizidine alkaloid monocrotaline. Thorax 38, 307-309.

LaFranconi, W.M., Duhamel, R.C., Brendel, K. and Huxtable, R.J. (1984) Differentiation of the cardiac and pulmonary effects of monocrotaline, a pyrrolizidine alkaloid. Biochem. Pharmacol. 33, 191-197.

LaFranconi, W.M., Ohkuma, S. and Huxtable, R.J. (1985) Biliary excretion of a novel pneumotoxic metabolite of the pyrrolizidine alkaloid, monocrotaline. Toxicon 23, 983-992.

Lame, M.W., Morin. D., Jones, A.D., Segall, H.J. and Wilson, D.W. (1990) Isolation and identification of a pyrrolic gluathione conjugate of the pyrrolizidine alkaloid monocrotaline. Toxicol. Lett. 51, 321-329.

Page 155: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

1 5 1

Lame, M.W., Jones, A.D., Morin, D., Segall, H.J. and Wilson, D.W. (1995) Biliary excretion of pyrrolic metabolites of ['"'Qmonocrotaline in the rat. Drug Metab. Disp. 23, 422-429.

Levy, H.R. (1979) Glucose-6-phosphate dehydrogenases. Adv. Enzymol. 48, 97-192.

Lenzen, S., Freytag, S., Panten, U., Flatt, P.R. and Bailey, C. (1990) Alloxan and ninhydrin inhibition of hexokinase from pancreatic islets and tumoural insulin-secreting cells. Pharmacol. Toxicol. 66, 157-162.

Logie, G.C, Grue, M.R., and Liddell, J.R. (1994) Proton NMR spectroscopy of pyrrolizidine alkaloids. Phytochemistry 37, 43-109.

Lyford, C.L., Vergava, G.G. and Moeller, D.D. (1976) Hepatic veno-occlusive disease originating in Ecuador. Gastroenterology 70, 105-108.

Marston, A. and Hostettmann, K. (1994) Counter-current chromatography as a preparative tool; applications and perspectives. J. Chromatogr. A 658, 315-341.

Mattocks, A.R. (1967) Spectrophotometric determination of unsaturated pyrrolizidine alkaloids Analyt. Chem. 39, 443-447.

Mattocks, A.R. (1968a) Toxicity of pyrrolizidine alkaloids. Nature 217, 723-728.

Mattocks, A. R. (1968b) Spectrophotometric determination of pyrrolizidine alkaloids -some improvements. Anal. Chem. 40, 1749-1750.

Mattocks, A.R. (1969) Dihydropyrrolizine derivatives from unsaturated pyrrolizidine alkaloids. J. Chem. Soc. C 8, 1155-1162.

Mattocks, A.R. (1971) The occurance and analysis of pyrrolizidine AZ-oxides. Xenobiotica 1, 451-453.

Mattocks, A.R. (1972) Acute hepatotoxicity and pyrrolic metabolites in rats dosed with pyrrolizidine alkaloids. Chem.-Biol. Interact. 5, 227-242.

Mattocks, A.R. (1981) Relation of structural features to pyrrolic metabolites in livers of rats given pyrrolizidine alkaloids. Chem.-Biol. Interact. 35, 301-310.

Mattocks, A.R. (1982) Hydrolysis and hepatotoxicity of retronecine diesters. Toxicol. Lett. 14, 111-116.

Mattocks, A.R. (1986) Chemistry and Toxicology of Pyrrolizidine Alkaloids. London; Academic Press.

Mattocks, A.R. and White, LN.H. (1970) Estimation of metabolites of pyrrolizidine alakloids in animal tissues. Anal. Biochem. 38, 529-535.

Mattocks, A.R. and White, LN.H. (1971) The conversion of pyrrolizidine alkloids to N-oxides and to dihydropyrrolizine derivatives by rat liver microsomes in vitro. Chem.-Biol. Interact. 3. 383-396.

Page 156: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

152

Mattocks, A.R. and White, I.N.H. (1973) Toxic effects and pyrrolic metabolites in the liver of young rats given the pyrrolizidine alkaloid retrorsine. Chern.-Biol. Iniemct. 6, 297-306.

Mattocks, A.R. and Legg, R.F. (1980) Antimitotic activity of dehydroretronecine, a pyrrolizidine alkaloid metabolite, and some analogous compounds, in a rat liver parenchymal cell line. Chem.-Biol. Interact. 30, 325-336.

Mattocks, A.R. and Bird, I. (1983a) Pyrrolic and A'-oxide metabolites formed from pyrrolizidine alkaloids by hepatic microsomes in vitro: relevance to in vivo hepatotoxicity. Chem.-Biol. Interact. 43, 209-222.

Mattocks, A.R. and Bird, I. (1983b) Alkylation by dehydroretronecine, a cytotoxic metabolite of some pyrrolizidine alkaloids: an in vitro test. Toxicol. Lett. 16, 1-8.

Mattocks, A.R., Jukes, R. and Brown, J. (1989) Simple procedures for preparing putative toxic metabolites of pyrrolizidine alkaloids. Toxicon 27, 561-567.

Mattocks, A.R., Legg, R.F. and R. Jukes, R. (1990) Trapping of short-lived electrophilic metabolites of pyrrolizidine alkaloids escaping from perfused rat liver. Toxicol. Lett. 54, 93-99.

Mattocks, A.R. and Jukes, R. (1990a) Recovery of the pyrrolic nucleus of pyrrolizidine alkaloid metabolites from sulphur conjugates in tissues and body fluids. Chem.-Biol. Interact. 75. 225-239.

Mattocks, A.R. and Jukes, R. (1990b) Trapping and measurement of short-lived alkylating agents in a recirculating flow system. Chem.-Biol. Interact. 76, 19-30.

Mattocks, A.R., Croswell, S., Jukes, R., and Huxtable, R.J. (1991) Identified of a biliary metabolite formed from monocrotaline in isolated, perfused rat liver. Toxicon 29,409-415.

Mattocks, A.R. and Jukes, R. (1992) Detection of sulphur-conjugated pyrrolic metabolites in blood and fresh or fixed liver tissue from rats given a variety of toxic pyrrolizidine alkaloids. Toxicol. Lett. 63, 47-55.

Mattocks, A.R. and Jukes, R. (1992) Chemistry of sulphur-bound pyrrolic metabolites in the blood of rats given different types of pyrrolizidine alkaloids. Nat. Toxins 1, 89-95.

Middleton, R.J. (1990) Hexokinases and glucokinases. Biochem. Soc. Trans. 18, 180-183.

Miranda, C.L., Reed, R.L., Guengerich, P.P., and Buhler, D.R. (1991). Role of Cytochrome P450IIIA4 in the metabolism of the pyrrolizidine alkaloid senecionine in the human liver. Carcinogenesis 12, 515-519.

Miranda, C.L., Chung, W., Reed, R.L., Zhao, X., Henderson, M.C., Wang, J., Williams, D.E. and Buhler, D.R. (1991) Havin-containing monooxygenase: a major

Page 157: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

153

detoxifying enzyme for the pynrolizidine alkaloid senecionine in guinea pig tissues. Biochem. Biophys. Res. Comm. 178, 546-552.

Miser, J.S., Smithson, W.A., Krivit, W., Hughes, C.H., Davis, D., Krailo, M.D. and Hammond, F.D. (1992) Phase II trial of indicine A'-oxide in relapsed acute leukemia of childhood. Am. J. Clin. Oncol. 15, 135-140.

Mohabbat, O., Srivastava, R.N., Younos, M.S., Sediq, G.G., Menzad, A.A. and Aram, G.N. (1976) An outbreak of hepatic veno-occlusive disease in north-western Afghanistan. Lancet 7, 269-271.

Munday, R., Ludwig, K. and Lenzen, S. (1993) The relationship between the physicochemical properties and the biological effects of alloxan and several A'-alkyl substituted alloxan derivatives. J. Endocrin. 139, 153-163.

Noltman, E.A., Gubler, C.J. and Kuby, S.A. (1961) Glucose-6-phosphate-dehydrogenase I. Isolation of the crystalline enzyme from yeast. J. Biol. Chem. 236, 1225-1230.

Ohtsubo, K., Ito, Y., Saito, M., Furuya, T. and Hikichi, M. (1977) Hypertrophy of pulmonary arteries and arterioles with cor pulmonale in rats induced by seneciphylline, a pyrrolizidine alkaloid. Experientia 33,498-499.

Pan, L.C., Wilson, D.W., Lame, M.W., Jones, A.D. and H.J. Segall. (1993) COR pulmonale is caused by monocrotaline and dehydromonocrotaline, but not by glutathione or cysteine conjugates of dihydropyrrolizine. Toxicol. Appl. Pharmacol. 118, 87-97.

Percy, J.J. and Pierce, A.E. (1971) Immunosuppresive activity of the pyrrolizidine alkaloid metabolite dehydroheliotridine. Immunology 21, 273-280.

Peterson, J.E. (1965) Effects of the pyrrolizidine alkaloid, lasiocarpine A^-oxide, on nuclear and cell division in the liver of rats. J. Path. Bacterol. 89, 153-175.

Peterson, J.E., Samuel, A. and Jago, M.V. (1972) Pathological effects of dehydroheliotridine, a metabolite of heliotrine based pyrrolizidine alkaloids in the young rat. J. Pathol. 107, 175-190.

Peterson, J.E. and Jago, M.V. (1980) Comparison of the toxic effects of dehydroheliotridine and heliotrine in pregnant rats and their embryos. J. Pathol. 131. 339-355.

Peterson, J.E. and Jago, M.V., Reddy, J.K. and Jarret, R.G. (1983) Neoplasia and chronic disease associated with the prolonged administration of dehydroheliotridine to rats. J. Natl. Cancer Inst. 70, 381-386.

Petry, T., Bowden, G.T., Huxtable, R.J. and Sipes, I. G. (1984) Characterization of hepatic DNA damage induced by the pyrrolizidine alkaloid monocrotaline. Cancer Res. 44, 1505-1509.

Page 158: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

154

Plestina, R. and Stoner, H.B. (1972) I^ilmonary oedema in rats given monocrotaline pyrrole. J. Pathol. 106, 235-249.

Puri, R.N., Bhatnagar, D., Roskoski Jr., R. (1988) Inactivation of yeast hexokinase by o-pthalaldehyde: evidence for the presence of a cysteine and a lysine at or near the active site. Biochim. Biophys. Acta 957, 34-46.

Puri, R.N. and Roskoski, R. (1994) Inactivation of yeast hexokinase by Cibacron Blue 3G-A: spectral, kinetic and structural investigations. Biochem. J. 300, 91-97.

Quails, C.W. and Segall, H.J. (1978) Rapid isolation and identification of pyrrolizidine alkaloids (Senecio vulgaris) by use of high performance liquid chromatography. J. Chomatogr. 150, 202-206.

Ratnoff, O.D. and Mirick, G.S. (1949) Influence of sex upon the lethal effects of the hepatotoxic alkaloid, monocrotaline. Bull. Johns Hopkins Hosp. 84, 507-525.

Reed, R.L., Miranda, C.L., Kedzierski, B., Henderson, M.C. and Buhler, D.R. (1992) Microsomal formation of a pyrrolic alcohol glutathione conjugate of the pyrrolizidine alkaloid senecionine. Xenobiotica 22, 1321-1327.

Ridker, P.M., Ohkumas, S., McDermott, W.V., Trey, C. and Huxtable, R.J. (1985) Hepatic veno-occlusive disease associated with the consumption of pyrrolizidine-containing dietary supplements. Gastroenterology 88, 1050-1054.

Robertson, K.A., Seymour, J.L., Hsia, M-T., and Allen, J.R. (1977) Covalent interaction of dehydroretronecine, a carcinogenic metabolite of the pyrrolizidine alkaloid monocrotaline. with cysteine and glutathione. Cancer Res. 37, 3141-3144.

Robertson, K.A. (1982) Alkylation of N2 in deoxyguanosine by dehydroretronecine, a carcinogenic metabolite of the pyrrolizidine alkaloid monocrotaline. Cancer Res. 42, 8-14.

Roeder, E., Liang, X.T. and Kabus, K.J. (1992) Pyrrolizidine alkaloids from the seeds of Crotalaria sessiliflora. Planta Med. 58, 283.

Roitman, J.N. (1983) The pyrrolizidine alkaloids of Amsinckia menziesii. Aust. J. Chem. 36, 769-778.

Roth, R.A., Dotzlaf, L.A., Baranyi, B. and Hook, J.B. (1981) Effect of monocrotaline ingestion on liver, kidney and lung of rats. Toxicol. Appl. Pharmacol. 60, 193-203.

Rothschild, M., Aplin, R.T., Cockrum, P.A., Edgar, J.A., Fairweather, P. and Lees, R. (1979) Pyrrolizidine alkloids in arctiid moths (Lep.) with a discussion on host plant relationships and the role of these secondary substances in the Arctiidae. Biol. J. Linn. Soc. 12, 305-326.

Samuel, A. and Jago, M.V. (1975) Localization in the cell cycle of the antimitotic action of the pyrrolizidine alkaloid, lasiocarpine, and of its metabolite, dehydroheliotridine. Chem.-Biol. Interact. 10, 185-197.

Page 159: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

155

Saville, B. (1962) Solvolysis of unsaturated sulfides assisted by silver ion. J. Chem. Soc. Perkin Trans. 4062-4068.

Schoental, R. (1959) Liver lesions in young rats suckled by mothers treated with the pyrrolizidine (Senecio) alkaloids, lasiocarpine and retrorsine. J. Pathol. Bacterial. 77. 485-504.

Schoental, R. (1970) Hepatotoxic activity of retrorsine, senkirkine and hydroxysenkirkine in newborn rats, and the role of epoxides in the carcinogenesis of pyrrolizidine alkaloids and aflatoxins. Nature (Lond.) 227, 401-402.

Schoental. R. (1968) Toxicology and carcinogenic action of pyrrolizidine alkaloids. Cancer Res. 28, 2237-2246.

Schultze, A.E., Wagner, J.G., White, S.M. and Roth, R.A. (1991) Early indications of monocrotaline pyrrole-induced lung injury in rats. Toxicol. Appl. Pharmacol. 109. 41-50.

Segall, H.J. (1984) Recent chromatogrpahic methods to isolate pyrrolizidine alkaloids. J. Liq. Chromatogr. 7, 377-392.

Shoham, M. and Steitz (1980) Crystallographic studies and model building of ATP at the active site of hexokinase. J. Mol. Biol. 140, 1-14.

Shtenberg, A.I. and Orlova, N.V. (1955) Etiology of the so-called Dzhalangarsk Encephalitis. Vopr. Pitaniya 14, 27-31.

Smimov, F.E., Grosheva, G.A. and Stoljarova, A.G. (1959a) Role of the plant Trichodesma incanum in the etiology of" suiljuk" disease of horses. Uzbek. Inst. Vet. Sbom. Nauch. Tr. 13, 173-185.

Smimov, F.E., Grosheva, G.A. and Stoljarova, A.G. (1959b) Action of the leaves and stems of Trichodesma incanum on cattle. Uzbek. Inst. Vet. Sbom. Nauch. Tr. 13, 195-198.

Smimov, F.E. and Stoljarova, A.G. (1959) Pathogenesis of Trichodesma poisoning (suiljuk) in horses. Uzbek. Inst. Vet. Sbom. Nauch. Tr. 13, 186-194.

Smith, L.W. and Culvenor, C.C.J. (1981) 7. Nat. Prod. 44, 121-152.

Sped, W., Stuppner, H., Gassner, I., Judmaier, W., Dietze, O. and Vogel, W. (1995) Reversible hepatic veno-occlusive disease in an infant after consumption of pyrrolizidine-containing herbal tea. Eur. J. Pediatr. 154, 112-116.

Stillman, A.E., Huxtable, R., Consroe, P., Kohnen, P. and Smith, S. (1977) Hepatic and veno-occlusive disease due to pyrrolizidine poisoning in A.rizona. Gastroenterology 73, 349-352.

Stuart, K.L. and Bras, G. (1956) Veno-occlusive disease of the liver in Barbados. West Indian Med. J. 5, 33-36.

Page 160: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

156

Stuart, K.L. and Bras, G. (1957) Veno-occlusive disease of the liver. Q. J. Med. 26, 291-315.

Tallarida, R.J. and Murray, R.B. (1986) Manual of Pharmacological Calculations with Computer Programs, pp. 121-125. New York: Springer.

Tandon, H.D., Tandon, B.N. and Mattocks, A.R. (1978) An epidemic of veno-occlusive disease in Afghanistan. Am. J. Gastroenerol. 72, 607-613.

Thorpe, E. and Ford. E.J.H. (1968) Developement of hepatic lesions in calves fed with ragwort {Senecio jacobaea). J. Camp. Pathol. 78, 195-205.

United Nations Environment Programme, International Labour Organization and World Health Organization. (1988) Environmental Health Criteria 80: Pyrrolizidine Alkaloids., pp. 1-345. Geneva: World Health Organization.

Valdivia, E., Lalich, J.J., Hayashi, Y. and Sonnad, J. (1967) Alterations in pulmonary alveoli after a single injection of monocrotaline. Arch. Pathol. 84, 64-76.

Vamvakas, S. and Anders, M.W. (1990) Formation of reactive intermediates by phase II enzymes: gluathione dependant bioactivation reactions. Biological Reactive Intermediates FV. Witmer. C.M. et al., eds. Plenum Press, New York, p. 13-24

White, I.N.H., Mattocks, A.R. and Butler, W.H. (1972) The conversion of the pyrrolizidine alkaloid retrorsine to pyrrolic derivatives in vivo and in vitro and its acute toxicity to various animal species. Chem.-Biol. Interact. 6, 207-218.

Wild, S.L. (1994) Pyrrolizidine Alkaloids: Hepatic Metabolism and Extrahepatic Toxicity. Ph.D. Dissertation, University of Arizona.

Williams, D.E. Reed. R.L.. Kedzeirski, B. Ziegler, D.M. and Buhler, D.R. (1989) The role of flavin-containing monooxygenase in the ^/-oxidation of the pyrrolizidine alkaloid senecionine. Drug Metab. Disp. 17, 380-392.

Wilson, J.E. (1995) Hexokinases. Rev. Physiol. Biochem. Pharmacol.

Witte, L., Rubiolo P.. Bicchi, C. and Hartman. T. (1993) Comparative analysis of pyrrolizidine alkaloids from nanaral sources by gas chromatogrpahy-mass spectrometry. Phytochemistry 32, 187-196.

Yan, C.C. and Huxtable, R.J. (1994) Quantitation of the hepatic release of metabolites of the pyrrolizidine alkaloid, monocrotaline. Toxicol. Appl. Pharmacol. 127, 58-63.

Yan, C.C.. Cooper, R.A., and Huxtable, R.J. (1995) The comparative metabolism of the four pyrrolizidine alakaloids. seneciphylline, retrorosine, monocrotaline, and trichodesmine in the isolated, perfused rat liver. Toxicol. Appl. Pharmacol. 133, 277-284.

Yunosov. S.Y. and Plekhanova, N.V. (1957a) The structure of incanine (in Russian). Doklady Akad. Nuak. Uzbek. S.S.R.5, 1 3 - 1 6 .

Page 161: Pyrrolizidine alkaloids: Chemical basis of toxicity...2 Results of Separation of Pyrrolizidine Alkaloid Fractions Using High-Speed Counter-Current Chromatography 64 3 Summary of Wet

157

Yunosov, S.Y. and Plekhanova, N.V. (1957b) The structure of trichodesmine (in Russian). Doklady Akad. Niiak. Uzbek. S.S.R. 6, 19-22.

Yunosov, S.Y. and Plekhanova, N.V. (1959) Alkaloids of Trichodesma incanum. Structure of incanine and trichodesmine. Zh. Obshchei Khimii 29, 677-684.

Zalkow, L.H., Asibal, C.F., Glinski, J.A., Bonetti, S.J., Gelbaum, L.T., VanDerveer. D. and Powis, G. (1988) Macrocyclic pyrrolizidine alkaloids from Senecio anonymus. Separation of a complex alkaloidal extract using droplet counter-current chromatography. J. Nat. Prod. 51, 690-702.

Zhang, T. (1984) Horizontal flow-through coil planet centrifuge: some practical applications of countercurrent chromatography. J. Chromatogr. 315, 287-297.

Zhang, T., Cai D. and Ito, Y. (1988) Separations of flavonoids and alkaloids in medicinal herbs by high-speed counter-current chromatography. J. Chromatogr. 435, 159-166.