Natural Toxins from Cyanobacteria (Blue-Green Algae) - American

cells or toxins from certain freshwater/brackish water species of Ana- baena, Aphanizomenon ... 0097-6156/90/0418-0087$06.00/0 ..... gram of lyophiliz...
1 downloads 0 Views 2MB Size
Chapter 6 Natural Toxins from Cyanobacteria (Blue-Green Algae) 1

Downloaded by NORTH CAROLINA STATE UNIV on January 4, 2013 | http://pubs.acs.org Publication Date: January 29, 1990 | doi: 10.1021/bk-1990-0418.ch006

Wayne W. Carmichael, Nik A. Mahmood , and Edward G. Hyde Department of Biological Sciences, Wright State University, Dayton, OH 45435

A c u t e lethal toxicity from cyanobacteria is caused by ingestion o f toxic cells o r toxins from certain freshwater/brackish water species o f Anabaena, Aphanizomenon, Microcystis, Nodularia, and Oscillatoria. C o n ­ tact poisonings are reported from marine species o f Lyngbya, Oscilla­ toria, and Schizothrix. M o r e recently, cytotoxic compounds with antialgal, antifungal, antibacterial, a n t i p r o t o z o a l and antineoplastic activity have been reported from species o f Scytonema, Oscillatoria, Hapalosiphon, and Plectonema. A c u t e l y lethal toxins include a related family o f hepatotoxic cyclic hepta and pentapeptides, termed microcystins o r cyanoginosins, which contain both D and L amino acids plus two novel amino acids. Species and strains o f Anabaena produce at least two neurotoxins, termed anatoxins; o n e a depolarizing neuromuscular blocking agent, the other an irreversible anticholinesterase. Strains o f Aphanizomenon flos-aquae produce saxitoxin and neosaxitoxin, the p r i ­ mary toxins i n cases o f paralytic shellfish poisoning ( P S P ) . These various toxins have intraperitoneal mouse LD values o f 10—500 μg/kg. Current research indicates that a third group o f toxins with contact irritant properties are produced by some freshwater cyanobac­ teria. Indirect evidence indicates that o n e o r a l l of these toxins are responsible for certain cases o f human gastroenteritis and dermatitis from m u n i c i p a l and recreational water supplies. 50

Reports o f toxic algae i n the freshwater environment are almost exclusively due to members o f the division Cyanophyta, c o m m o n l y called blue-green algae o r cyanobacteria. A l t h o u g h cyanobacteria are found i n almost any environment rang­ ing from h o t springs to A n t a r c t i c soils, k n o w n toxic members are mostly planktonic. Published accounts o f field poisonings by cyanobacteria are k n o w n since the late 19th century (95). These reports describe sickness and death o f livestock, pets, and wildlife following ingestion o f water containing toxic algae cells o r the toxin released by the aging cells. Recent reviews o f these poisonings a n d the toxins o f freshwater cyanobacteria are given by Carmichael (1—3), C o d d and B e l l (4), and G o r h a m and Carmichael (5).

Current address: Springborn Laboratory, Inc., Mammalian Toxicology Division, 553 North Broadway, Spencerville, OH 45887 0097-6156/90/0418-0087$06.00/0 o 1990 American Chemical Society

In Marine Toxins; Hall, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1990.

MARINE TOXINS: ORIGIN, STRUCTURE, AND MOLECULAR PHARMACOLOGY

Downloaded by NORTH CAROLINA STATE UNIV on January 4, 2013 | http://pubs.acs.org Publication Date: January 29, 1990 | doi: 10.1021/bk-1990-0418.ch006

88

Toxins produced by cyanobacteria are i n two general groups. T h e first, which is emphasized i n this chapter, includes those toxins responsible for acute lethal poisonings. A b o u t 12 genera have been implicated i n producing these toxins but only Anabaena, Aphamzomenon, Microcystis, Nodularia, and Oscillatoria have h a d toxins isolated, and investigated chemically and toxicologically. T h e second group includes a number o f secondary chemicals that are n o t highly lethal to animals but instead show m o r e selective bioactivity. These bioactive chemicals (Table I) include scytophycins produced by certain strains o f Scytonema pseudohofmanni. S. pseudohofmanni strain B C - 1 - 2 , which was isolated from a forested area o n the island o f O a h u , H a w a i i , was found to produce two l i p o p h i l i c toxins termed scytophycin A ( M W 821) and B ( M W 819). Scytophycin B is moderately toxic t o mice by the intraperitoneal route ( L D = 650 fig/kg). Scytophycins show a very strong cytotoxic activity, however, when tested against cell cultures, such as K B h u m a n epiderm o i d carcinoma and N I H / 3 T 3 mouse fibroblast cell lines (1 and 0.65 ng/mL, respectively). These toxins are also moderately active against intraperitoneally implanted P388 lymphocytic leukemia and Lewis lung carcinoma (6). S. hofinanni U T E X 1581 has been shown to produce the chlorine-containing diaryl-lactone called cyanobacter i n . H i i s c o m p o u n d has been shown to have anticyanobacterial activity and has been proposed as a possible algicide against cyanobacteria (7,8). A cytotoxic alkaloid has been isolated from the filamentous species Hapalosiphon fontinalis strain V - 3 - 1 . This isolate was made from soil samples collected i n the M a r s h a l l Islands i n 1981. This strain produces the l i p o p h i l i c c o m p o u n d hapalindole A , which has a broad range o f antialgal and antimycotic activity (9). Oscillatoria acutissima strain B - l , isolated from a freshwater p o n d i n O a h u , was found to produce two novel macrolide compounds termed acutiphycin and 20, 2 1 didehydroacutiphycin. These macrolides show cytotoxicity ( K B a n d N I H / 3 T 3 ) and antitumor activity ( M u r i n e Lewis l u n g carcinoma) (10). O t h e r toxins that show l o w lethal toxicity to laboratory test animals include lipopolysaccharide endotoxin produced as part o f the cell wall by a l l cyanobacteria (11) a n d certain toxins o f some cyanobacteria suspected o f causing contact irritation i n recreational water supplies (4,12; Carmichael and C o d d , unpublished results). E c o n o m i c losses due to water-based diseases o f freshwater cyanobacteria toxins have o n l y been reported from the first group and are the result o f contact with o r c o n s u m p t i o n o f water containing toxin and/or toxic cells. These toxins are a l l water soluble and temperature stable. They are either released by the cyanobacterial cell o r loosely b o u n d so that changes i n cell permeability o r age allow their release into the environment. K n o w n occurrences o f toxic cyanobacteria i n water supplies include Canada (four provinces), E u r o p e (12 countries), U n i t e d States (20 states), U S S R (Ukraine), Australia, India, Bangladesh, S o u t h Africa, Israel, Japan, N e w Z e a l a n d , Argentina, C h i l e , Thailand, and the Peoples R e p u b l i c o f C h i n a (13-15). T h e economic impact from toxic freshwater cyanobacteria include the costs incurred from deaths o f domestic animals; allergic and gastrointestinal problems after h u m a n contact with waterblooms (including lost i n c o m e from recreational areas); a n d increased expense for the detection and removal o f taste, odor, and toxins. T h e remainder o f this chapter discusses the acute lethal toxins produced by freshwater cyanobacteria which includes hepatotoxins and neurotoxins (Table I). 5 0

Neurotoxins Anatoxins. Neurotoxins produced by filamentous Anabaena flos-aquae are called anatoxins ( A N T X S ) (15). Currently two anatoxins, from different strains o f A. flos-aquae, have been isolated and at least partially characterized. A N T X - A from strain (single filament isolate) N R C - 4 4 - 1 is the first toxin from a freshwater

In Marine Toxins; Hall, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1990.

6.

CARMICHAEL ET AL.

Natural Toxins from Cyanobacteria

89

Table I. Toxins of Freshwater Cyanobacteria Species, Strain, and Source

Toxin Term

Structure

LD ug/kg ip, Mouse

Anatoxin-A

Secondary amine alkaloid, M W 165

200

50

Neurotoxins

Downloaded by NORTH CAROLINA STATE UNIV on January 4, 2013 | http://pubs.acs.org Publication Date: January 29, 1990 | doi: 10.1021/bk-1990-0418.ch006

A. flos-aquae Strain N R C - 4 4 - 1 (Canada, Saskatchewan) Strain N R C - 5 2 5 - 1 7 (Canada, Saskatchewan)

Aph. flos-aquae Strain N H - 1 & N H - 5 (U.S., N e w Hampshire)

Anatoxin-A(S)

JV-hydroxy guanidine methyl-phosphate ester, M W 252

20

Aphantoxin (neosaxitoxin)

P u r i n e alkaloid M W 315 ( n e o S T X ) M W 299 ( S T X )

10

Heptapeptides M W 994

50

Heptapeptides M W 909-1044

50

A p h a n t o x i n II (saxitoxin)

Hepatotoxins A. flos-aquae Strain S-23-g-l (Canada, Saskatchewan)

Microcystes

3

Cyanoginosins M. aeruginosa Strain W R - 7 0 ( = U V - 0 1 0 ) (South Africa, Transvaal)

3

(Waterbloom, Australia, N e w S o u t h Wales)

Cyanoginosin

Heptapeptide M W 1035

50

(Waterbloom, U . S . , Wisconsin

Microcystin

Heptapeptide M W 994

50

Strain N R C - 1 ( S S - 1 7 ) (Canada, O n t a r i o )

Microcystin

Heptapeptide M W 994

50

Strain 7820 Microcystin (Scotland, L o c h Balgaves)

Heptapeptide M W 994

50

(Waterbloom, Norway, L a k e Akersvatn)

Microcystin

Heptapeptide M W 994

50

Microcystin

Heptapeptide M W 994 M W 1044

50

M. aeruginosa Strain M - 2 2 8 (Japan, T o k y o )

See text for explanation o f terminology. Continued on next page

In Marine Toxins; Hall, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1990.

MARINE TOXINS: ORIGIN, STRUCTURE, AND MOLECULAR PHARMACOLOGY

Table I. Species, Strain, and Source

Continued LD ug/kg ip, Mouse 5Q

Structure

Toxin Term

M. aeruginosa

Downloaded by NORTH CAROLINA STATE UNIV on January 4, 2013 | http://pubs.acs.org Publication Date: January 29, 1990 | doi: 10.1021/bk-1990-0418.ch006

M W 1039 M. viridis

Cyanoviridin

N. spumigena

Nodularin

3

Heptapeptide M W 1039

not reported

Pentapeptide M W 824

30-50

O. agardhii var. isothrix Microcystins (Waterbloom, Norway, L a k e Froylandsvatn)

Heptapeptides M W 1009

300-500

O. agardhii var. Microcystins (Waterbloom, Norway, L a k e Kolbotnvatn)

Heptapeptides M W 1023

500-1000

Cytotoxins S. pseudohofmanni Strain B C - 1 - 2 (U.S., Hawaii)

Scytophycin

Methylform amide A & B A = M W 821; B = M W 819

650 (scytophycin B )

S. hofinanni Strain U T E X - 1 5 8 1 (U.S., Texas)

Cyanobacterin

Chlorinated diaryllactone

not reported

H. fontinalis Strain V - 3 - 1 (Marshall Islands)

Hapalindole A

Substituted indole alkaloid

not reported

T. byssoidea Strain H - 6 - 2 (U.S., H a w a i i )

Tubercidin

Pyrrolopyrimidine

not reported

O. acutissima Strain B - l (U.S., Hawaii)

Acutiphycin

Macrolide

not reported

3

See text for explanation o f terminology.

In Marine Toxins; Hall, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1990.

Downloaded by NORTH CAROLINA STATE UNIV on January 4, 2013 | http://pubs.acs.org Publication Date: January 29, 1990 | doi: 10.1021/bk-1990-0418.ch006

6.

CARMICHAEL ET AL.

Natural Toxins from Cyanobacteria

91

cyanobacteria to be chemically defined. It is the secondary amine, 2-acetyl-9azabicyclo[4.2.1]non-2-ene (16,17) molecular weight 165 daltons (Figure 1). It has been synthesized through a ring expansion o f cocaine (18,19), from i m i n i u m salts (20-22), from nitrone (23,24), from 4-cycloheptenone o r tetrabromotricyclooctane (25) by construction o f the azabicyclo ring from 9-methyl-9-azabicyclo[3.3.1]nonanl - o l (26), and by starting with 9-methyl-9-aza[4.2.1]nonan-2-one (27). A N T X - A is a potent, postsynaptic, depolarizing, neuromuscular blocking agent that affects both nicotinic and muscarinic acetylcholine ( A C H ) receptors (28—31). Signs o f poisoning i n field reports for wild and domestic animals include staggering, muscle fasciculations, gasping, convulsions, and opisthotonos (birds). D e a t h by respiratory arrest occurs w i t h i n minutes to a few hours depending o n species, dosage, and prior food consumption. T h e L D intraperitoneal (ip) mouse for purified toxin is about 200 /xg/kg body weight, with survival time o f 4 - 7 m i n . Thus, animals need to ingest only a few milliliters to a few liters o f the toxic surface b l o o m to receive a lethal bolus (32—34). Detection o f A N T X - A , w h i l e still primarily by mouse bioassay, is being supplemented by three analytical detection methods. These methods are based o n high performance liquid chromatography ( H P L C ) (35,36), gas chromatography-mass spectrometry ( G C - M S ) (37, H i m b e r g personal communication), and gas chromatography-electron capture detection ( G C E C D ) (38). 5 Q

A l l k n o w n occurrences o f A N T X - A production have been from Canada o r the U n i t e d States. M o r e recently, A N T X - A has been detected i n A. flos-aquae blooms from Japan (Watanabe, personal communication), Norway (Skulberg, personal c o m munication), and F i n l a n d (99). A neurotoxin more recently isolated and under current study is referred to as anatoxin-a(s) [ A N T X - A ( S ) ] . T h e primary source o f this toxin is A. flos-aquae strain N R C - 5 2 5 - 1 7 isolated i n 1965 from Buffalo P o u n d L a k e i n Saskatchewan, Canada. A N T X - A ( S ) is physiologically and chemically different from A N T X - A It produces opisthotonos i n chicks, as does A N T X - A , but also causes viscous salivation [which gives the terminology its (S) label] and lachrymation i n mice, c h r o m o dacryorrhea (bloody tears) i n rats, urinary incontinence, muscular weakness, fasciculation, convulsion and defecation prior to death by respiratory arrest. Also observed is a dose-dependent fasciculation o f limbs for 1—2 m i n after death. A N T X - A ( S ) was purified, from A. flos-aquae cells, by c o l u m n chromatography and H P L C (39,40), and its structure is given i n Figure 1 (100). A N T X - A ( S ) is acid stable, unstable i n basic conditions, has very l o w U V absorbance, gives a positive alkaloid test, and has a molecular weight o f 253 ( m + H ) daltons. The L D i p mouse for A N T X - A ( S ) is about 20 ug/kg, ten times more lethally toxic than A N T X - A A t the L D the survival time for mice is 10—30 m i n . M a h m o o d and Carmichael (41) concluded from the signs o f poisoning, i.e., salivat i o n and muscle twitch potentiation, that the cholinergic system was the primary target o f A N T X - A ( S ) . N o serum cholinesterase activity was observed i n rats dosed w i t h 350 and 600 ug/kg o f A N T X - A ( S ) , suggesting an anticholinesterase mechanism. A N T X - A ( S ) also sensitized the frog rectus abdominus muscle and chick biventer cervicis muscle to exogenous acetylcholine, but had no effect o n the action o f A N T X - A (a depolaring agent) o n frog rectus abdominus. Further w o r k with A N T X - A ( S ) (42) has confirmed and extended the conclusion that this toxin is an anticholinesterase agent. A N T X - A ( S ) was shown to inhibit i n vitro electric eel acetylcholinesterase ( A C h E , E C 3.1.1.7) and horse serum butyrylcholinesterase ( B U C h E , E C 3.1.1.8) i n a time and concentration dependent fashion by a mechanism similar to the organophosphate anticholinesterases as illustrated i n Scheme I. +

5 Q

5 0

In Marine Toxins; Hall, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1990.

Downloaded by NORTH CAROLINA STATE UNIV on January 4, 2013 | http://pubs.acs.org Publication Date: January 29, 1990 | doi: 10.1021/bk-1990-0418.ch006

MARINE TOXINS: ORIGIN, STRUCTURE, AND MOLECULAR PHARMACOLOGY

R=H; saxitoxin dihydrochloride R=OH; neosaxitoxin dihydrochloride Figure 1. Left: A n a t o x i n - a ( A N T X - A ) hydrochloride. Produced by the freshwater filamentous cyanobacterium Anabaena flos-aquae N R C - 4 4 - 1 . R i g h t : Anatoxin-a(s). Produced by the freshwater filamentous cyanobacterium Anabaena flos-aquae N R C - 5 2 5 - 1 7 . Bottom: Aphantoxin-I (neosaxitoxin) and A p h a n t o x i n - I I (saxitoxin) produced by certain strains o f the filamentous cyanobacterium Aphanizomenon flos-aquae.

In Marine Toxins; Hall, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1990.

6.

Natural Toxins from Cyanobacteria

CARMICHAEL E T A L .

EOH+IX

j ±

EOH(IX)

—•

k_,

EOI

EOH+H+IO"

HX

93

(I)

H 0 2

In this scheme, E O H is the enzyme, I X is the i n h i b i t o r (either a carbamate o r an organophosphate). E O H ( I X ) is analogous to the Michaelis M e n t o n complex seen w i t h the substrate reaction. E O I is the acyl-enzyme intermediate for carbamates o r a phosphoro-enzyme intermediate for the organophosphates. T h e equilib r i u m constant for this reaction ( K ) is defined as kj/k j and the phosphorylation or carbamylation constant is defined as 1^. In this study (42), A N T X - A ( S ) was found to be more specific for A C h E than B U C h E . T h e double reciprocal and D i x o n plot o f the inhibition o f electric eel A C h E indicated that the toxin is a non-competitive i n h i b i t o r ( V decreases, k remains unchanged) (Figure 2). Since non-competitive inhibitors can be reversible o r irreversible, a plot o f V vs. the total amount o f enzyme was used to distinguish which o f the two A N T A - A ( S ) resembled. In this type o f plot the lines for c o n t r o l and non-competitive i n h i b i t o r w i l l pass through the origin, while that o f the irreversible i n h i b i t o r parallels the control. Figure 3 shows that A N T X - A ( S ) is an irreversible i n h i b i t o r o f electric eel AChE. Further comparison o f the kinetic result o f A N T X - A ( S ) i n h i b i t i o n o f A C h E with diisopropyl-fluorophosphate ( D F P ; an irreversible cholinesterase i n h i b i tor) indicated that i n h i b i t i o n follows the generalized scheme for irreversible i n h i b i tors and that the phosphorylation constant (2 per min) was five times less than D F P (10 per m i n ) . T h e affinity constant o f D F P (55.2 /xg/mL) is also m u c h weaker (~ 110 fold) than A N T X - A ( S ) (0.5 ug/mh). T h e bimolecular i n h i b i t i o n constant (Kj = k ^ ^ indicates that A N T X - A ( S ) is 22 times m o r e powerful than DFP.

Downloaded by NORTH CAROLINA STATE UNIV on January 4, 2013 | http://pubs.acs.org Publication Date: January 29, 1990 | doi: 10.1021/bk-1990-0418.ch006

d

m a x

m

M o r e recent unpublished w o r k is continuing o n the molecular mechanism o f A C h E / A N T X - A ( S ) binding. This w o r k focuses o n two areas: the first is c o m parison o f the kinetic inhibition for different sources o f A C h E by A N T X - A ( S ) and then comparing the kinetic constants against that found with D F P . T h e second set o f investigations is l o o k i n g at the nature o f the interaction o f A N T X - A ( S ) and A C h E . B i n d i n g o f substrate and the irreversible inhibitors o f A C h E involves attachment at two sites o n the A C h E molecule (43,44). These two sites are 5 A apart and comprise the active site o f the enzyme. T h e first is k n o w n as the anionic site which contains one o r more negatively charged residues surrounded by regions o f hydrophobicity (45). This area binds the cationic p o r t i o n o f the molecule, and for uncharged molecules the hydrophobic regions act as an anchoring point (46). T h e second site is the esteratic site where the ester linkage o f the substrate or inhibitor is cleaved by a serine residue made highly reactive by a "charge-relay" system made up o f a histidine and an aspartate o r tyrosine residue. There are two ways o f producing inhibition by anticholinesterase agents. T h e first, termed irreversible inhibition, occurs when a carbamate o r organophosphate reacts with A C h E . T h e serine residue can become carbamylated with the subsequent hydrolysis o f the carbamyl group. This process requires only minutes for reactivation o f the enzyme. In the second process, the serine becomes phosphorylated and it requires hours to days for hydrolysis back to the active enzyme. T h e second type o f inhibition, termed reversible, includes true reversible anticholinesterases (i.e., tetramethylammonium, T M A ) which block o n l y the anionic site and are displaced by increasing substrate concentration. A C h E can also be i n h i bited by high substrate concentrations where the cationic head o f one molecule and the ester linkage o f another molecule b i n d concurrently at the active site and substrate hydrolysis cannot occur resulting i n inhibition (44,47). In o u r studies to date, and reported here, we have used A C h , 3,3'-dimethylbutylacetate (an uncharged

In Marine Toxins; Hall, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1990.

94

MARINE TOXINS: ORIGIN, STRUCTURE, AND MOLECULAR

Downloaded by NORTH CAROLINA STATE UNIV on January 4, 2013 | http://pubs.acs.org Publication Date: January 29, 1990 | doi: 10.1021/bk-1990-0418.ch006

3

PHARMACOLOGY

r-

Figure 2. I n h i b i t i o n o f eel A C h E by A N T X - A ( S ) - the 'secondary plot'. P , the firstorder rate constant which was the rate o f inhibition at that A N T X - A ( S ) concentration obtained from the 'primary p l o t ' (insert). T h e intercept o n the 1/P axis is l / k ^ and the intercept o n the 1/[I] axis is - 1 / K . Figure insert: Progressive irreversible i n h i b i t i o n o f eel A C h E by A N T X - A ( S ) . T h e inactivation followed first-order kinetics. A N T X - A ( S ) concentrations, /zg/mL: ( A ) 0.083; ( • ) 0.166; (o) 0.331; (•) 0.497; ( V ) 0.599; ( • ) control. E a c h point represents the mean o f 3 or 4 determinations. d

In Marine Toxins; Hall, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1990.

6.

Natural Toxins from Cyanobacteria

CARMICHAEL ET AL.

95

carbon isostere o f acetylcholine), and T M A to investigate the i n h i b i t i o n mechanism of A N T X - A ( S ) . In the first group o f studies, involving kinetic i n h i b i t i o n studies, comparisons o f the equilibrium ( K ) , phosphorylation ( K ~ ) , and i n h i b i t i o n constant (Kj) for the i n h i b i t i o n o f electric eel and human erythrocyte A C h E by A N T X - A ( S ) and D F P were done (Table II). F r o m Table II it is seen that A N T X - A ( S ) has a higher affinity for h u m a n erythrocyte A C h E ( K = 0 . 2 5 3 uM) than electric eel A C h E ( K = 3 . 6 7 uM). A N T X - A ( S ) also shows greater affinity for A C h E than D F P ( K = 3 0 0 uM). A n d finally the bimolecular rate constant, K j , which indicates the overall rate o f reaction, shows A C h E is more sensitive toward i n h i b i t i o n by A N T X A ( S ) ( K p l . 3 6 uM' min ) than D F P (K = 0.033 / i M ^ m i n ) . These studies add information to the comparative activity o f A N T X - A ( S ) and other irreversible A C h E inhibitors but do not show the site o f inhibition. The second group o f studies was designed to investigate the site o f i n h i b i t i o n . The results o f these protection studies are shown i n Figures 4 — 7. Figure 4 is the dose-activity curve to various concentrations o f A N T X - A ( S ) after a 2-min preincubation. Figures 5, 6, and 7 show the same dose-activity curves i n the presence o f 0.33 m M A C h , 2.6 m M 3,3'-dimethyl butylacetate and 0.11 m M T M A C a l c u l a t i o n o f a protection index (PI) (slope o f % activity/min i n presence o f protectant divided by the slope o f % activity/min i n absence o f protectant; P I < 1 , protection; >1 n o protection) shows that all compounds protect A C h E from A N T X - A ( S ) i n h i b i t i o n . Referenced to the 0.1 /zg A N T X - A ( S ) curve, the protective indices are: A C h , 0.5501; D M B A , 0.6460; and T M A 0.8598. F r o m this o n e can see that T M A protects only slightly, indicating that A N T X - A ( S ) site o f attachment is not primarily the anionic site. A C h and D M B A both protect the enzyme although A C h is slightly better than D M B A Protection by D M B A indicates that A N T X - A ( S ) does not inhibit by the substrate inhibition mechanism. In addition to these laboratory studies o n A N T X - A ( S ) , field poisonings by A N T X - A ( S ) have n o w been confirmed in at least o n e case (48). T h e poisonings involved the death o f five dogs, eight pups, and two calves that ingested quantities o f A. flos-aquae, containing A N T X A ( S ) , i n R i c h m o n d Lake, S o u t h D a k o t a , i n late summer 1985. In summary, A N T X - A ( S ) uses the two-site attachment mechanism analogous to substrate and does not inhibit A C h E i n the manner o f the reversible anticholinesterases. d

d

d

d

l

l

- 1

Downloaded by NORTH CAROLINA STATE UNIV on January 4, 2013 | http://pubs.acs.org Publication Date: January 29, 1990 | doi: 10.1021/bk-1990-0418.ch006

{

A p h a n t o x i n s . Occurrence o f neurotoxins (aphantoxins) i n the freshwater filamentous cyanobacterium Aphanizomenon flos-aquae was first demonstrated by Sawyer et al. (49). A l l aphantoxins ( A P H T X S ) studied to date have come from waterblooms and laboratory strains o f nonfasciculate (non-flake forming) Aph. flos-aquae that occurred i n lakes and ponds o f N e w H a m p s h i r e from 1966 through 1980. Toxic cells and extracts o f Aph. flos-aquae were shown to be toxic to mice, fish, and waterfleas (Daphnia catawba) by J a k i m and G e n t i l e (50). Chromatographic and pharmacological evidence established that A P H T X S consist mainly o f two neurotoxic alkaloids that strongly resemble saxitoxin ( S T X ) and neosaxitoxin ( n e o S T X ) , the two primary toxins o f red tide paralytic shellfish poisoning ( P S P ) (57). T h e b l o o m material and toxic strain used i n studies before 1980 came from collections made between 1960 and 1970. T h e more recent work o n A P H T X S has used two strains ( N H - 1 and N H - 5 ) isolated by Carmichael i n 1980 from a small p o n d near D u r h a m , N e w H a m p s h i r e (52,55). These A P H T X S , like n e o S T X and S T X , are fast-acting neurotoxins that inhibit nerve conduction by blocking sodium channels without affecting permeability to potassium, the transmembrane resting potential, o r membrane resistance (54). M a h m o o d and Carmichael (55), using the N H - 5 strain, showed that batch-cultured cells have a mouse i p L D o f about 5 mg/kg. E a c h 5 Q

In Marine Toxins; Hall, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1990.

96

MARINE TOXINS: ORIGIN, STRUCTURE, AND MOLECULAR PHARMACOLOGY

Downloaded by NORTH CAROLINA STATE UNIV on January 4, 2013 | http://pubs.acs.org Publication Date: January 29, 1990 | doi: 10.1021/bk-1990-0418.ch006

5 r

Ej (units) Figure 3. P l o t o f V against total enzyme [ET] showing the irreversible i n h i b i t i o n o f electric eel acetylcholinesterase ( A C h E ) by A N T X A ( S ) . T h e enzymes were incubated with 0.32 / i g / m L A N T X - A ( S ) for 1.0 m i n and acetylthiocholine (final concentrations 2.5, 4.7, 6.3, and 4

7.8 x 10" M ) was added. V was determined from the double reciprocal plots (not shown). K e y : (o) control; (•) A N T X - A ( S ) . (Reproduced w i t h permission from Ref. 42. Copyright 1987 Pergamon Press)

Table II. Affinity Equilibrium ( K ) , Phopshorylation Rate ( k p and Biomolecular Rate (1c,) Constants for the Inhibition of AChE by ANTX-A(S) and DFP d

AChE Electric Eel

Human Erythrocytes +

Toxin

(A*M)

(min )

OiM^min" )

K * d OiM)

ANTX-A(S)

3.67

2.0

1.36

0.25

21.40

84

DFP

300

10

0.03

_

_

_

k

(min" )

i (pM^min" )

K

K

d

1

1

and 1^ were determined from the secondary plots (Figure 2). K is the reciprocal o f the abscissa intercept and d

1

the reciprocal o f the

ordinate intercept. +

1

A n estimated M W o f 267 for A N T X - A ( S ) was used to calculate K . d

In Marine Toxins; Hall, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1990.

Downloaded by NORTH CAROLINA STATE UNIV on January 4, 2013 | http://pubs.acs.org Publication Date: January 29, 1990 | doi: 10.1021/bk-1990-0418.ch006

CARMICHAEL ET AL.

Natural Toxins from Cyanobacteria

Time (minutes) Figure 4. Concentration-dependent i n h i b i t i o n o f electric eel A C h E by A N T X - A ( S ) . A C h E and A N T X - A ( S ) were incubated for 2 m i n . before i n h i b i t i o n rate was determined. K e y : ( A ) 0.079 /xg/mL; ( • ) 0.032 /xg/mL; ( V ) 0.016 /xg/mL; «>) 0.0032 uglmU (•) -0016 / x g / m L N O T E : Total i n h i b i t i o n occurs when 0.158 /xg/mL A N T X - A ( S ) is preincubated with A C h E for two minutes.

Figure 5. T h e effects o f 0.11 m M acetylcholine o n the concentrationdependent i n h i b i t i o n o f A C h E by A N T X - A ( S ) . A c e t y l c h o l i n e was coincubated w i t h A N T X - A ( S ) and A C h E two minutes before the i n h i b i t i o n rate was determined. K e y : (o) 0.158 /xg/mL; ( A ) 0.079 /xg/mL; ( • ) 0.032 /xg/mL; ( V ) 0.016 jxg/mL; (0) 0.0032 /xg/mL; (•) .0016 uglmL. NOTE: T o t a l i n h i b i t i o n occurs when 0.158 ug/mL A N T X - A ( S ) is preincubated w i t h A C h E for two minutes.

In Marine Toxins; Hall, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1990.

Downloaded by NORTH CAROLINA STATE UNIV on January 4, 2013 | http://pubs.acs.org Publication Date: January 29, 1990 | doi: 10.1021/bk-1990-0418.ch006

MARINE TOXINS: ORIGIN, STRUCTURE, AND MOLECULAR PHARMACOLOGY

Time (minutes) Figure 6. T h e effects o f 2.6 m M 3,3'-dimethylbutyl acetate o n the concentration-dependent i n h i b i t i o n o f A C h E by A N T X - A ( S ) . Conditions and symbols are the same as Figure 5.

5 I 0

1

1

1

1

1

1

2

3

4

5

Time (minutes) F i g u r e 7. T h e effect o f 0.11 m M tetramethyl a m m o n i u m iodide o n the concentration dependent inhibition o f A C h E by A N T X - A ( S ) . C o n d i t i o n s and symbols are the same as Figure 5.

In Marine Toxins; Hall, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1990.

6.

CARMICHAEL ET AL.

Natural Toxins from Cyanobacteria

99

Downloaded by NORTH CAROLINA STATE UNIV on January 4, 2013 | http://pubs.acs.org Publication Date: January 29, 1990 | doi: 10.1021/bk-1990-0418.ch006

gram o f lyophilized cells yields about 1.3 m g aphantoxin I (neosaxitoxin) a n d 0.1 mg aphantoxin II (Saxitoxin) (Figure 1). A l s o detected are three labile neurotoxins that are not similar to any o f the k n o w n paralytic shellfish poisons. S h i m i z u et a l . (56) studied the biosynthesis o f the S T X analog n e o S T X using Aph. flos-aquae N H - 1 . They were able to confirm its presence i n strain N H - 1 and to explain the biosynthetic pathway for this important group o f secondary chemicals. Hepatotoxins. Low-molecular-weight peptide toxins that affect the liver have been the predominant toxins involved i n cases o f animal poisonings d u e to cyanobacterial toxins (2,5,57). After almost 25 years o f structure analysis o n toxic peptides o f the c o l o n i a l bloom-forming cyanobacterium Microcystis aeruginosa, Botes et a l . (58—60) and Santikarn et a l . (61) provided structure details o n o n e o f four toxins (designated toxin B E - 4 ) produced by the South African M. aeruginosa strain W R 7 0 ( = U V - 0 1 0 ) . They concluded that it was monocyclic and contained three D - a m i n o acids (alanine, erythro-^-methylaspartic acid, and glutamic acid), two L - a m i n o acids (leucine and alanine), plus two unusual amino acids. These were Nmethyldehydroalanine (Medha) and a nonpolar side chain o f 20 carbon atoms that turned o u t to be a novel 0-amino acid; 3-amino-O-methoxy-2,6,8-trimethyl-10phenyldeca-4,6-dienoic acid ( A D D A ) . Based o n fast atom bombardment mass spectrometry ( F A B M S ) and nuclear magnetic resonance ( N M R ) studies, B E - 4 toxin is n o w k n o w n to be a cyclic heptapeptide having a molecular weight o f 909 daltons (Figure 8). Instead o f calling the B E - 4 toxin microcystin, as previous Microcystis toxins were called (62-64) and using alphabetical o r numerical suffixes to indicate chromatographic elution order o r structural differences, Botes (60) proposed the generically derived designation cyanoginosin ( C Y G S N ) . This prefix, w h i c h indicates that cyanobacterial origin, is followed by a two-letter suffix that indicates the identity and sequence o f the two L - a m i n o acids relative to the N - M e - d e h y d r o a l a n y l - D alanine bond. Thus toxin B E - 4 was renamed c y a n o g i n o s i n - L A since leucine and alanine are the L - a m i n o acids. These L - a m i n o acids were shown by E l o f f et a l . (65) a n d Botes et a l . (66) to vary between strains and to account for structural differences for different toxic fractions w i t h i n a single strain. Botes et a l . (66) showed that the other three toxins o f strain W R - 7 0 a l l h a d the same three D amino acids and two novel amino acids ( M e d h a and A D D A ) . T h e l - a m i n o acids were leucine- arginine ( C Y G S N - L R ) , tyrosine-arginine ( C Y G S N - Y R ) , a n d tyrosinealanine ( C Y G S N - Y A ) . They were also able to show that the hepatotoxin isolated by E l l e m a n et a l . (67) from waterbloom material collected i n M a l p a s D a m , N e w S o u t h Wales, Australia, contained the five characteristic amino acids plus the L r amino acid variants tyrosine-methionine (therefore it is termed C Y G S N - Y M ) . Microcystin ( M C Y S T ) is the term given to the fast death factor ( F D F ) produced by M. aeruginosa strain N R C - 1 and its daughter strain N R C - 1 (SS-17) (62,68). A n absolute structure for the toxin o f strain N R C - 1 (SS-17) is not yet available but is k n o w n to be a peptide ( M W 994) containing the variant amino acids leucine and arginine (Carmichael, unpublished). Krishnamurthy et a l . (69,70) have shown that the toxin isolated from a waterbloom o f M. aeruginosa collected i n L a k e Akersvatn, Norway (77), has a structure similar to that o f M Y C S T from N R C - 1 (SS-17) and C Y G S N - L R . This toxin has also been found to be the main toxin produced by the Scottish strain o f M. aeruginosa P C C - 7 8 2 0 and a Canadian A. flos-aquae strain S23-g-l (69,70) (Figure 8). T h e identification o f a peptide toxin from A. flos-aquae S-23-g-l provides the first evidence that these hepatotoxins are produced by filamentous as well as coccoid cyanobacteria. A. flos-aquae S-23-g-l and toxic M. aeruginosa from a waterbloom i n Wisconsin also produced a second cyclic

In Marine Toxins; Hall, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1990.

Downloaded by NORTH CAROLINA STATE UNIV on January 4, 2013 | http://pubs.acs.org Publication Date: January 29, 1990 | doi: 10.1021/bk-1990-0418.ch006

100

MARINE TOXINS: ORIGIN, STRUCTURE, AND MOLECULAR PHARMACOLOGY

Figure 8. Left: T h e cyclic heptapeptide hepatotoxin m i c r o c y s t i n - L A (cyanoginosin-LA) produced by the colonial cyanobacterium Microcystis aeruginosa strain W R - 7 0 ( U V - 0 1 0 ) . M W = 909. R i g h t : T h e cyclic heptapeptide hepatotoxin microcystin-LR (cyanoginosin-LR) produced by a waterbloom o f the colonial cyanobacterium Microcystis aeruginosa collected i n L a k e Akersvatn, Norway, 1984-85; M W = 9 9 4 , (69,71).

In Marine Toxins; Hall, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1990.

Downloaded by NORTH CAROLINA STATE UNIV on January 4, 2013 | http://pubs.acs.org Publication Date: January 29, 1990 | doi: 10.1021/bk-1990-0418.ch006

6.

CARMICHAEL ET AL.

Natural Toxins from Cyanobacteria

101

heptapeptide hepatotoxin, which has been found to have six o f the same amino acids, that is, leucine-arginine, but has aspartic acid instead o f £ - m e t h y l a s p a r t i c acid (69). T h e filamentous genus Oscillatoria has also been shown t o produce a hepatotoxin (72,73). F r o m water blooms o f O. agardhii var. and O. agardhii var. isothrix, two similar cyclic heptapeptides have been isolated. B o t h toxins have the variant amino acids arginine-arginine and aspartic acid instead o f ^-methylaspartic acid. O. agardhii var. isothrix also has dehydroalanine instead o f methyldehydroalanine (70). M o r e recently M. viridis (74) and M. aeruginosa (75) have been shown to produce the cyclic heptapeptide with an arginine-arginine " L " amino acid variant. Nodularia spumigena has also been shown to produce a peptide w i t h hepatotoxic activity. T h e more recent reports come from Australia (76), the G e r m a n D e m o c r a t i c R e p u b l i c (77), D e n m a r k (78), Sweden (79), a n d F i n l a n d (80,81). Recently structure information o n Nodularia toxin has been presented by R i n e h a r t (97) for waterbloom material collected i n L a k e Forsythe, N e w Zealand, i n 1984; by Eriksson et a l . (81) from waterbloom material collected i n the Baltic Sea i n 1986, and Runnegar et a l . (82) for a field isolate from the Peel Inlet, Perth, Australia. Structure w o r k by Rinehart, Eriksson, and Runnegar a l l indicate that the peptide is smaller than the heptapeptide toxins. Rinehart's w o r k (97) indicates the toxin is a pentapeptide with a similar structure to the heptapeptides a n d containing fimethylaspartic acid, glutamic acid, arginine, dehydrobutyrine, and A D D A ( M W 824). T h e hepatotoxins have been called Fast-Death Factor (68), microcystin (62), cyanoginosin (60), cyanoviridin (74), and cyanogenosin (apparently a misspelling o f cyanoginosin) (75). Because o f the different terms used to refer to similar c o m pounds it has been proposed (83) that the term microcystin ( M C Y S T ) plus the suffix " X Y " (designating the variant L - a m i n o acids) be used as the basis for naming a l l the existing a n d future monocyclic heptapeptide toxins o f cyanobacteria. Cyclic peptides with fewer o r greater than seven peptides o r peptide linked components (i.e., the pentapeptide from Nodularia is termed nodularin [ N O D L N ] ) should be named according to the genus name from which they are originally isolated o r by their chemical composition relative to the existing microcystins. T h e hepatotoxic heptapeptide general structure thus becomes:

1 2 3 tyclo(-D-Ala-L- X -D-eiythro-0-methy^ M

4

5

6

7

H

X = L e u c i n e ( L ) , A r g i n i n e ( R ) , Tyrosine ( Y ) Y = Arginine (R), Alanine (A), Methionine (M) " X Y " combinations for heptapeptide toxins currently defined: L R ; L A ; Y A ; Y M ; YR; R R A D D A = 3-amino-9-methoxy-2,6,8-trimethyl-10-phenyldeca-4,6-dienoic acid

M o d e o f A c t i o n for Microcystins. T h e liver has always been reported as the organ that showed the greatest degree o f histopathological change when animals are poisoned by these cyclic peptides. T h e molecular basis o f action for these cyclic peptides is n o t yet understood but the cause o f death from toxin and toxic cells administered to laboratory mice and rats is at least partially k n o w n a n d is c o n cluded to be hypovolemic shock caused by interstitial hemorrhage into the liver (84). This w o r k w i t h small animal models is currently being extended to larger

In Marine Toxins; Hall, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1990.

102

MARINE TOXINS: ORIGIN, STRUCTURE, AND MOLECULAR PHARMACOLOGY

animals i n order to study the uptake, distribution, and metabolism o f the toxins (85). There is evidence to show from studies using I-labeled M C Y S T - Y M ( C Y G S N - Y M ) that the liver is the organ for b o t h accumulation and excretion (86,87). B r o o k s a n d C o d d (88), using C labeled M C Y S T - L R , showed that 7 0 % o f the labeled toxin was localized i n the mouse liver after 1 m i n following i p injection o f the toxin. Studies at b o t h the light and electron microscopic ( E M ) level o f time-course histopathological changes i n mouse liver show rapid and extensive centrilobular necrosis o f the liver w i t h loss o f characteristic architecture o f the hepatic cords. Sinusoid endothelial cells and then hepatocytes show extensive fragmentation and vesiculation o f cell membranes (89,90). U s i n g m i c r o c y s t i n - L R from M. aeruginosa strain P C C - 7 8 2 0 , D a b h o l k a r and Carmichael (97) found that at b o t h lethal and sublethal toxin levels hepatocytes show progressive intracellular changes beginning at about 10 m i n postinfection. T h e most c o m m o n response to lethal and sublethal injections is vesiculation o f rough endoplasmic reticulum ( R E R ) , swollen m i t o c h o n dria, and degranulation (partial o r total loss o f ribosomes from vesicles). T h e vesicles appear to form from dilated parts o f R E R by fragmentation o r separation. Affected hepatocytes remain intact and do not lyse. U s e o f the isolated perfused rat liver to study the pathology o f these toxins shows similar results to the i n vivo work. B e r g et a l . (92) used three structurally different cyclic heptapeptide hepatotoxins ( M C Y S T - L R , desmethyl 3 - M C Y S T - R R , and didesmethyl 3 , 7 - M C Y S T - R R ) . A l l three toxins had a similar effect o n the perfused liver system although b o t h " R R " toxins required higher concentrations (5-7 x ) to produce their effect. This finding was consistent w i t h the lower toxicity o f the " R R toxins, which was about 500 a n d 1000 /ig/kg i p mouse compared to 50 /ig/kg for M C Y S T - L R . 125

Downloaded by NORTH CAROLINA STATE UNIV on January 4, 2013 | http://pubs.acs.org Publication Date: January 29, 1990 | doi: 10.1021/bk-1990-0418.ch006

1

4

M

In vitro studies o n isolated cells including hepatocytes, erythrocytes, fibroblasts, and alveolar cells continue to demonstrate the specificity o f action that these toxins have for liver cells (83,86,93). This specificity has led A u n e and Berg (94) to use isolated rat hepatocytes as a screen for detecting hepatotoxic waterblooms o f cyanobacteria. T h e cellular/molecular mechanism o f action for these cyclic peptide toxins is n o w an area o f active research i n several laboratories. These peptides cause striking ultrastructural changes i n isolated hepatocytes (95) including a decrease i n the polymerization o f actin. This effect o n the cells cytoskeletal system continues to be investigated and recent w o r k indirectly supports the idea that these toxins interact w i t h the cells cytoskeletal system (86,96). W h y there is a specificity o f these toxins for liver cells is n o t clear although it has been suggested that the bile uptake system may be at least partly responsible for penetration o f the toxin into the cell (92).

Summary A c u t e poisoning o f humans by freshwater cyanobacteria as occurs with paralytic shellfish poisoning, w h i l e reported, has never been confirmed. H u m a n s are probably just as susceptible as pets, livestock, o r wildlife but people naturally avoid c o n tact w i t h heavy waterblooms o f cyanobacteria. I n addition, there are n o k n o w n vectors, like shellfish, to concentrate toxins from cyanobacteria into the human food chain. Susceptibility o f humans to cyanobacteria toxins is supported mostly by indirect evidence. I n many o f these cases, however, i f a more thorough epidemiological study had been possible these cases probably w o u l d have shown direct evidence for toxicity. In addition to acute lethal poisonings, episodes o f dermatitis and/or irritation

In Marine Toxins; Hall, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1990.

Downloaded by NORTH CAROLINA STATE UNIV on January 4, 2013 | http://pubs.acs.org Publication Date: January 29, 1990 | doi: 10.1021/bk-1990-0418.ch006

6. CARMICHAEL ET AL.

Natural Toxins from Cyanobacteria

103

from contact with freshwater cyanobacteria are occurring with increasing frequency. This is in part because more eutrophic waters are being used for recreational purposes. Investigations in the United States (72), Canada (14), Scotland (4), and Norway (73) have shown that dermatotoxic blooms may be dominated by Anabaena, Aphanizomenon, Gloeotrichia, and Oscillatoria, respectively. In summary, it can be reported that toxic cyanobacteria can produce neurotoxic, hepatotoxic, and dermatotoxic compounds that are a direct threat to animal and human water supplies. This threat increases as water bodies become more eutrophic, thus supporting higher production of toxic and nontoxic cyanobacteria. Presence of these potent natural product toxins poses an increasing threat to the maintenance of quality water supplies for agriculture, municipal, and recreational use. Acknowledgments Work with European toxic cyanobacteria was partially supported by a N A T O collaborative research grant between W.W. Carmichael and G . A . Codd, University of Dundee, Scotland, and O . M . Skulberg, Norwegian Water Research Institute, Oslo, Norway. Toxin structure work on European and North American peptide toxins is supported in part by U.S. A M R D C contract DAMD17-87-C-7019 to W.W. Carmichael. Portions of the work represent part of the Ph.D. dissertation research of N . A Mahmood and E . G . Hyde. Their work was supported in part by fellowship support from the Biomedical Ph.D. Program, Wright State University. Literature Cited 1. Carmichael, W. W. In The Water Environment—Algal Toxins and Health; Carmichael, W. W., Ed.; Plenum: New York, 1981; p 1. 2. Carmichael, W. W. In Advances in Botanical Research; Callow, E . A., Ed.; Academic: London, 1986; Vol. 12, p 47. 3. Carmichael, W. W. In Handbook of Natural Toxins, Marine Toxins and Venoms; Tu, A. T., Ed.; Marcel Dekkar: New York, 1988; Vol. 3, p 121. 4. Codd, G. A.; Bell, S. G. Water Pollution Control 1985, 84, 225-32. 5. Gorham, P. R.; Carmichael, W. W. In Algae and Human Affairs; Lembi, C. A.; Waaland, J. R., Eds.; Cambridge Univ. Press: Oxford, 1988; p 403. 6. Moore, R. E.; Patterson, G. M. L.; Mynderse, J. L.; Barchi, J., Jr. Pure Appl. Chem. 1986, 58, 263-71. 7. Mason, C. P.; Edwards, K. R.; Carlson, R. E., Pignatello, J.; Glenson, F. R.; Wood, J. M. Science 1982, 215, 400-02. 8. Gleason, F. K.; Paulson, J. L. Arch. Microbiol. 1984, 138, 273-77. 9. Moore, R. E.; Cheuk, C.; Patterson, G. M. L. J. Am. Chem. Soc. 1984, 106, 6456-57. 10. Barchi, J. J., Jr.; Moore, R. E.; Patterson, G. M. L. J. Am. Chem. Soc. 1984, 106, 8193-97. 11. Raziuddin, S.; Siegelman, H. W.; Tornabene, T. G. Eur. J. Biochem. 1983, 137, 333-36. 12. Billings, W. H . In The Water Environment: Algal Toxins and Health; Carmichael, W. W., Ed.; Plenum: New York, 1981; p 243. 13. Skulberg, O. M.; Codd, G. A.; Carmichael, W. W. Ambio 1984, 13, 244-47. 14. Carmichael, W.W.; Jones, C. L. A.; Mahmood, N. A.; Theiss, W. W. In Critical Reviews in Environmental Control; Straub, C. P., Ed.; CRC Press: Florida, 1985; p 275. 15. Carmichael, W. W.; Gorham, P. R. Mitt. Int. Verein. Limnol. 1978, 21, 285-95. 16. Huber, C. S. Acta Crystallograph. 1972, B28, 2577-82.

In Marine Toxins; Hall, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1990.

104 17. 18. 19. 20. 21.

Downloaded by NORTH CAROLINA STATE UNIV on January 4, 2013 | http://pubs.acs.org Publication Date: January 29, 1990 | doi: 10.1021/bk-1990-0418.ch006

22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39.

40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53.

MARINE TOXINS: ORIGIN, STRUCTURE, AND MOLECULAR PHARMACOLOGY Devlin, J. P.; Edwards O. E.; Gorham, P. R.; Hunter, N. R.; Pike, P. K.; Stavric, B. Can. J. Chem. 1977, 55, 1367-71. Campbell, H . F.; Edwards, O. E., Kolt, R. J. Can. J. Chem. 1977, 55, 1372-79. Campbell, H . F.; Edwards, O. E.; Edler, J. W., Kolt, R. J. Pol. J. Chem. 1979, 53, 27-37. Bates, H. A.; Rapoport, H. J. Am. Chem. Soc. 1979, 101, 1259-65. Peterson, J. S.; Toteberg-Kaulen, S.; Rapoport, H . J. Org. Chem. 1984, 49, 2948-53. Koskinen, M. P.; Rapoport, H. J. Med. Chem. 1985, 28, 1301-09. Tufariello, J. J.; Mechler, H.; Senaratne, K. P. A. J. Am. Chem. Soc. 1984, 106, 7979-80. Tufariello, J. J.; Meckler, H.; Senaratne, K. P. A. Tetrahedron 1985, 41, 3447-53. Danheiser, R. L.; Morin, J. M., Jr.; Salaski, E. J. J. Am. Chem. Soc. 1985, 107, 8066-73. Wiseman, J. R.; Lee, S. Y. J. Org. Chem. 1986, 51, 2485-87. Lindgren, B.; Stjernlof, P.; Trozen, L. Acta ChemicaScand.1987, B41, 180-83. Carmichael, W. W.; Biggs, D. F.; Peterson, M. A. Toxicon 1979, 17, 229-36. Spivak, C. E.; Witkop, B.; Albuquerque, E. X. Mol. Pharm. 1980, 18, 384-94. Spivak, C. E.; Waters, J.; Witkop, B.; Albuquerque, E. X. Mol. Pharm. 193, 23, 337-43. Aronstam, R. S.; Witkop, B. Proc. Natl. Acad. Sci. USA 1981, 78, 4639-43. Carmichael, W. W.; Gorham, P. R. J. Phycol. 1977, 13, 97-101. Carmichael, W. W.; Gorham, P. R.; Biggs, D. F. Can. Vet. J. 1977, 18, 71-5. Carmichael, W. W.; Biggs, D. F. Can. J. Zool. 1978, 56, 510-12. Astrachan, N. B.; Archer, B. G. In The Water Environment: Algal Toxins and Health; Carmichael, W. W., Ed.; Plenum: New York, 1981; p 437. Wong, S. H.; Hindin, E. Am. Water Works Assoc. J. 1982, 74, 528-29. Smith, R. A.; Lewis, D. Vet. Hum. Toxicol. 1987, 29, 153-4. Stevens, D. K.; Krieger, R. I. J. Analytical Tox. 1988, 12, 126-31. Carmichael, W. W.; Mahmood, N. A. In Seafood Toxins; Ragelis, E . P., Ed.; American Chemical Soc. Symposium Series 262, Washington D.C., 1984; p 377. Mahmood, N. A. Ph.D. Thesis, Wright State University, Dayton, Ohio, 1985. Mahmood, N. A.; Carmichael, W. W. Toxicon 1986, 24, 425-34. Mahmood, N. A.; Carmichael, W. W. Toxicon 1987, 25, 1221-27. Rosenberry, T. In Advances in Enzymology, Meister, A., Ed.; John Wiley and Sons: New York, 1975; p 103. Main, A. R. In Introduction to Biochemical Toxicology; Hodgson, E.; Guthrie, F., Eds.; Elsevier: New York, 1980; p 193. Wilson, I. B.; Quan, C. Arch. Biochem. Biophys. 1958, 73, 131-43. Husan, F. B.; Cohen, J. B. J. Biol. Chem. 1980, 255, 3896-3904. Tomlinson, G.; Mutusi, B.; McLennan, I. Mol. Pharm.. 1980, 18, 33-9. Mahmood, N. A.; Carmichael, W. W.; Pfahler, D. Am. J. Vet. Res. 1988, 49, 500-03. Sawyer, P. J.; Gentile, J. H.; Sasner, J. J., Jr. Can. J. Microbiol. 1988, 14, 1199-204. Jakim, E.; Gentile, J. H. Science 1968, 162, 915-16. Sasner, J. J., Jr.; Ikawa, M.; Foxall, T. L. In Seafood Toxins; Ragelis, E . P., Ed.; American Chem. Soc. Symp. Series: Washington D.C., 1984; p 391. Carmichael, W. W. S. Afr. J. Sci. 1982, 78, 367-72. Ikawa, M.; Wegner, K.; Foxall, T. L.; Sasner, J. J., Jr. Toxicon 1982, 20, 747-52.

In Marine Toxins; Hall, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1990.

Downloaded by NORTH CAROLINA STATE UNIV on January 4, 2013 | http://pubs.acs.org Publication Date: January 29, 1990 | doi: 10.1021/bk-1990-0418.ch006

6. CARMICHAEL ETAL.

Natural ToxinsfromCyanobacteria

105

54. Adelman, W. J., Jr.; Fohlmeister, J. F.; Sasner, J. J., Jr.; Ikawa, M. Toxicon 1982, 20, 513-16. 55. Mahmood, N. A.; Carmichael, W. W. Toxicon 1986, 24, 175-86. 56. Shimizu, Y.; Norte, M.; Hori, A.; Genenah, A.; Kobayashi, M. J. Am. Chem. Soc. 1984, 106, 6433-34. 57. Schwimmer, M.; Schwimmer, D. In Algae, Man and the Environment; Jackson, D. F., Ed.; Syracuse Univ.: Syracuse, 1968; p 279. 58. Botes, D. P.; Druger, H.; Viljoen, C. C. Toxicon 1982, 20, 945-54. 59. Botes, D. P.; Viljoen, C. C., Kruger, H.; Wessels, P. L.; Williams, D. H . J. Chem. Soc. Perkin. Trans. 1982, 1, 2742-48. 60. Botes, D. P. In Mycotoxins and Phycotoxins; Steyn, P. S.; Vleggaar, R., Eds.; Elsevier: Amsterdam, 1986; Vol. 1, p 16. 61. Santikarn, S.; Williams, D. H.; Smith, R. J.; Hammond, S. J.; Botes, D. P. J. Chem. Soc. Chem. Commun. 1983, 12, 652-54. 62. Konst, H.; McKercher, P. D.; Gorham, P. R.; Robertson, A.; Howell, J. Can. J. Comp. Med. Vet. 1965, 29, 221-28. 63. Murthy, J. R.; Capindale, J. B. Can. J. Biochem. 1970, 48, 508-10. 64. Rabin, P.; Darbre, A. Biochem. Soc. Trans. 1975, 3, 428-30. 65. Eloff, J. N. Ph.D. Thesis; Univ. of the Orange Free State, R.S.A., 1987; 8 chapters. 66. Botes, D. P.; Wessels, H.; Kruger, H.; Runnegar, M. T. C.; Santikarn, S.; Smith, R. J.; Barna, J. C. J.; Williams, D. H. J. Chem. Soc. Perkin, Trans. 1985, 1, 2741-48. 67. Elleman, T. C.; Falconer, I. R.; Jackson, A. R. B.; Runnegar, M . T. Aust. J. Biol. Sci. 1978, 31, 209. 68. Bishop, C. T.; Anet, E. F. L. J.; Gorham, P. R. Can. J. Biochem. Physiol. 1959, 37, 453-71. 69. Krishnamurthy, T.; Carmichael, W. W.; Sarver, E. W. Toxicon 1986, 24, 865-73. 70. Krishnamurthy, T.; Szafraniec, L.; Sarver, E. W.; Hunt, D. F.; Shabanowitz, J.; Carmichael, W. W.; Missler, S.; Skulberg, O.; Codd, G. Proc. Am. Soc. Mass Spec. (ASMS) - Cincinnati, OH, 1986; p 93. 71. Berg, K.; Carmichael, W. W.; Skulberg, O. M.; Benestad, C.; Underrdal, B. Hydrobiologia 1987, 144, 97-103. 72. Ostensvik, O.; Skulberg, O. M.; Soli, N. E . In The Water Environment: Algal Toxins and Health; Carmichael, W. W., Ed.; Plenum: New York, 1981; p 315. 73. Eriksson, J.; Meriluoto, J. A. O.; Kujari, H. P.; Skulberg, O. M . Comp. Biochem. Physiol. 1988, 89, 207-10. 74. Kusumi, T.; Oui, T.; Watanabe, M. M.; Takahogh, H.; Kakisawa, H. Tetrahed. Letters 1987, 28, 4695-98. 75. Painuly, P.; Perez, R.; Fukai, T.; Shimizu, Y. Tetrahed. Letters1988,29, 11-14. 76. Main, D. C.; Berry, P. H.; Peet, R. L.; Robertson, J. P. Aust. Vet. J. 1977, 53, 578-81. 77. Kalbe, L.; Tiess, D. Arch. Exp. Vet. Med. 1964, 18, 535-39. 78. Lindstrom, E. Dansk. Vet. Tidsskr. 1976, 59, 637-4 79. Edler, L.; Ferno, S.; Lind, M. G.; Lundberg, R.; Nilsson, P. O. Ophelia 1985, 24, 103-09. 80. Persson, P. E.; Sivonen, K.; Keto, J.; Kononen, K.; Niemi, M.; Viljamaa, H . Aqua Fenn. 1984, 14, 147-54. 81. Eriksson, J. E.; Meriluoto, J. A. O.; Kujari, H. P.; Osterlund, K.; Fagerlund, K.; Hallbom, L. Toxicon 1988, 26, 161-66. 82. Runnegar, M . T. C.; Jackson, A. R. B.; Falconer, I. R. Toxicon 1988, 26, 143-5.

In Marine Toxins; Hall, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1990.

Downloaded by NORTH CAROLINA STATE UNIV on January 4, 2013 | http://pubs.acs.org Publication Date: January 29, 1990 | doi: 10.1021/bk-1990-0418.ch006

106

MARINE TOXINS: ORIGIN, STRUCTURE, AND MOLECULAR PHARMACOLOGY

83. Carmichael, W. W.; Beasley, V. A.; Bunner, D. L.; Eloff, J. N.; Falconer, I.; Gorham, P. R.; Harada, K-I.; Yu, M-J.; Krishnamurthy, T.; Moore, R. E.; Rinehart, K.; Runnegar, M.; Skulberg, O. M.; Watanabe, M. Toxicon 1988, 26, 971-3. 84. Theiss, W. C.; Carmichael, W. W.; Wyman, J.; Brunner, R. Toxicon 1988, 26, 603-13. 85. Beasley, V. R.; Lovell, R.; Cook, W.; Lunden, G.; Holmes, K.; Hooser, S.; Haschek-Hock, W.; Carmichael, W. W. Proc. Am. Chem. Soc., 8th Rocky Mt. Regional Meeting, 1986 (Abstract). 86. Falconer, I. R.; Runnegar, M. T. C. Chem.Biol.Interactions 1987, 63, 215-25. 87. Runnegar, M. T. C.; Falconer, I. R. Toxicon 1987, 24, 105-15. 88. Brooks, W. P.; Codd, G. A. Pharm. Tox. 1987, 60, 187-91. 89. Runnegar, M. T. C.; Falconer, I. R. In The Water Environment: Algal Toxins and Health; Carmichael, W. W., Ed.; Plenum: New York, 1981; p 325. 90. Foxall, T. L.; Sasner, J. J., Jr. In The Water Environment: Algal Toxins and Health; Carmichael, W. W., Ed.; Plenum: New York, 1981; p 365. 91. Dabholkar, A. S.; Carmichael, W. W. Toxicon 1987, 25, 285-92. 92. Berg, K.; Wyman, J.; Carmichael, W. W.; Dabholkar, A. S. Toxicon 1988, 26, 827-37. 93. Runnegar, M. T. C.; Andrews, J.; Gerdes, R. G.; Falconer, I. R. Toxicon 1987, 25, 1235-39. 94. Aune, T.; Berg, K. J. Toxicol Environ. Health 1986, 19, 325-36. 95. Runnegar, M. T. C.; Falconer, I. R. Toxicon 1986, 24, 105-15. 96. Eriksson, J.; Hagerster, H.; Isomaa, B. Biochem. Biophysc. Acta 1987, 930, 304-10. 97. Rinehart, K.; Harada, K.-I; Namikoshi, M.; Chen, C.; Harvis, C.A.; Munroe, M.H.G.; Blunt, J.W.; Mulligan, P.E.; Beasley, U.R.; Dahmen, A.M.; Carmichael, W.W. J. Am. Chem. Soc. 1988, 110, 8557-8. 98. Francis, G. Nature (London) 1878, 18, 11-2. 99. Sivonen, K.; Himberg, K.; Luukkamen, R.; Niemelä, S.I.; Poon, G.K.; Codd, G.A. Tox. Assess., 1989, 4, 339-52. 100. Matsunaga, S.; Moore, R. E.; Niemczurd, W. P.; Carmichael, W. W.J.Am. Chem. Soc. 1989, 111, 8021-23. RECEIVED August 9, 1989

In Marine Toxins; Hall, S., et al.; ACS Symposium Series; American Chemical Society: Washington, DC, 1990.