• Nem Talált Eredményt

Monoamine and Polyamine Analogues

N/A
N/A
Protected

Academic year: 2022

Ossza meg "Monoamine and Polyamine Analogues"

Copied!
26
0
0

Teljes szövegt

(1)

M o n o a m i n e a n d Polyamine Analogues

E. A. Zeller

I. General Remarks 53 A. Chemical Reactivity of Amines 53

B. Enzymology of Amine Metabolism 54

II. Amine Analogues 59 A. Polyamines and Mescaline 59

B. α-Alkylamines 6 1

C. Phenylcyclopropylamines ( P C P ) 63

D. Hydrazine Derivatives 65

III. Concluding Remarks 7 3

References ^4

I. GENERAL REMARKS

A. Chemical Reactivity of Amines

An attempt is made in this brief chapter to describe the biological response elicited by some amine analogues in terms of their interaction with certain enzymes. Only aliphatic and arylalkylamines, but not aro­

matic amines, amino acids, or amino sugars, are considered.

In aqueous media amines are capable of acting on many parts of enzyme systems. Primary as well as secondary bonds may be established as the following selection indicates: (a) Amino groups are able to share an electron pair with nucleophilic residues and tend to initiate the forma­

tion of Sehiff bases and metal complexes; they are also apt to participate in transaminase reactions (Section I, Β, 1). Alkylation of the amino group eliminates some of these processes, e.g., the production of Schiff bases.

(6) Amino compounds are capable of taking part in the formation of hydrogen bonds, either as hydrogen donators or receptors. In every en­

zyme system many opportunities for hydrogen-bonding arise. The helical

5 3

(2)

5 4 Ε. Α. ZELLER structure of proteins and deoxyribonucleic acids testifies to the efficiency of hydrogen bonds in which nitrogen is involved. Obviously, quaternary amines are excluded from this type of interaction, (c) According to Grimm et al (1), the isosteric residues — Ο — , — N H — , and — C H2— display such chemical and physical properties that sensitive physical and biologi­

cal criteria and tests often fail to differentiate among them {2). If C H2 residues adjacent to terminal amino groups are replaced by the isosteric N H groups, hydrazine derivatives are obtained. We expect, therefore, to find many similarities in the behavior of hydrazines and of their parent amines (Section II, D ) .

Clearly, amine analogues may and do affect enzymic systems in more than one way. If there are two or more amine residues present in one molecule, they do not necessarily affect the enzyme moiety in an identical manner. An example of the multiplicity of reaction types is given for diamine oxidase (Section I, B, 3).

Obviously, the non-nitrogen moiety of the molecule may also contribute to the binding forces between amines and enzymes. Van der Waals' forces should be mentioned in general, and the many different kinds of inter­

action known to exist between aromatic rings in particular. A ring-to-ring attachment has been suggested to play a role in the binding of substrates and inhibitor molecules to the active site of monoamine oxidase (Sec­

tion I, B, 2).

Many drugs contain aliphatic side chains ending with dialkylated amino residues. Chlorpromazine and imipramine are two examples. Since the alkylamine forms only a relatively small part of the molecules and since some of these compounds are treated in other chapters (e.g., chapter on tranquilizers), they are omitted here.

B. Enzymology of Amine Metabolism

Mono- and polyamines are involved in many enzymic processes; to describe them all would go far beyond the allotted space and the com­

petence of this author. The reactions include (a) the removal of the amino residue either by hydrolysis, by dismutation, by oxidative deamination, or by transamination [transamination between histamine and pyruvic acid has been observed recently (3)], (b) alkylation and dealkylation, (c) oxidation to form nitroso compounds [the nitrosamine derivative of imipramine was found in urine ( 4 ) ] , and (d) conversion to amides, e.g., peptides and acetylated amines. Since the biological roles of some of the processes listed are insufficiently known, they are left out here, as are the acetylation processes which are conventionally treated as a

(3)

part of the field of CoA enzymes. Chemical changes that occur outside the amine residues, e.g., hydroxylation of the β-carbon of dopamine, O-methylation of catechol amines, or glucosiduronide formation, are not presented here.

Thus, we are left with the discussion of three enzymes: transgluta­

minase, monoamine oxidase, and diamine oxidase. Only a few points will be raised which will be required as minimal background for our discus­

sion. Many equally, or even more, important data unfortunately have to be left out.

1. T R A N S G L U T A M I N A S E

In the years 1 9 5 4 - 1 9 5 7 Block (5) and Sarkar et al. (6) observed that radioactively labeled amines, such as phenylethylamine and cadaverine, were incorporated into soluble proteins by covalent bonds after being incu­

bated with liver extracts. Further analysis of this phenomenon by Waelsch and his co-workers led to the characterization of the enzyme transgluta­

minase (7, 8); under its influence and in the presence of C a + + a number of mono- and diamines replaced the nitrogen of glutamine residues

[Eq. ( 1 ) ] .

RCONH2 + H2NR' -> RCONHR' + N H3 (1) A number of proteins, e.g., pepsin, α-globulin, and fibrinogen, serve as

amine acceptors. [In insulin, the most effective protein, both glutamine residues of chain A are replaceable by cadaverine moieties.]

We may expect that the physicochemical and biological properties of the products in Eq. ( 1 ) differ markedly from those of the starting proteins.

2 . M O N O A M I N E O X I D A S E S

The primary step catalyzed by monoamine oxidase (MAO) is sum­

marized in Eq. ( 2 ) (9-1 la).

R C H 2 N H 2 + 02 + H20 -> RCHO -1- N H3 + H202 (2) The enzyme is widely distributed throughout the animal organism and

is found in cells derived from all three germinal layers. Intracellularly, it is located in mitochondria {9-11) and in lysosomes (12). Organs such as the brain (IS, 14) and the kidney (15) display specific MAO distribution patterns. Judging from statistical data, the activity level in a given organ and species fluctuates amazingly little. In various parts of the human brain the standard deviations amounted to less than 1 0 % (13) and in

(4)

56 Ε. Α. ZELLER mouse liver to less than 13% (15a). This consistency, however, does not mean that the enzyme levels are rigidly fixed throughout the whole life­

span. On the contrary, MAO changes markedly during ontogenesis (13, 16) and sexual maturation and responds strongly to the administration of steroid hormones (15a). From these data we receive the impression that the enzyme activity is regulated by an efficient and flexible feedback mechanism which adjusts the MAO level to the need of the amine metabolism.

There exist differences in MAO activity, not only among different organs of one species, but also among the same organs of different species.

These variations are not entirely of a quantitative, but also of a qualita­

tive nature. As compared with tyramine, the standard substrate, serotonin is attacked to a greater extent by brain MAO than by the liver oxidase (17), and irans-2-phenylcyclopropylamine inhibits brain enzyme in vitro and in vivo significantly better than liver MAO (Section II, C). In com­

paring extensively the liver oxidases of rabbit and of cattle, we are led to the conclusion that the active sites of the two enzymes differ sharply from one another (18, 19). MAO, therefore, is not a single entity, but comprises a whole family of closely related homologous enzymes.

In order to evaluate the possibility that a given amine or amine ana­

logue can act on MAO, we have to know something about the mechanism of enzyme action and about the architecture of the active site. A con­

siderable amount of information is summarized in Eq. (2). The general formula of a MAO substrate (disregarding iV-methylation) is shown in (I). Many varied observations are compatible with the assumption of

R C H 2 N H 2

(I)

an initial removal of one of the two α-hydrogens (19-21). In this process the substrate molecule may well be bound by covalent bond to the enzyme molecule. In the Eqs. (3-5), En and Ac are the symbols for the enzyme molecule and the hydrogen acceptor, respectively.

R C H 2 N H 2 + EnH + Ac RCHNH2 + AcH2 (3) En

I

RCHNH2 + H20 -> RCHO + N H3 + EnH (4)

I

En

AcH2 + 02 Ac + H2O2 (5)

(5)

Equations ( 3 - 5 ) give a more detailed interpretation of the process of oxidative deamination than Eq. ( 1 ) .

The residue R of formula (I) can be substituted by aromatic, arylalkyl, and aliphatic groups, whereby in the latter case the optimal length of the aliphatic chain varies from species to species as seen in the classic paper by Alles and Heegard (22). As has been shown in this laboratory, the substitution of the ring of arylalkylamines by nitro, methyl, chloro, amino, and other residues deeply affects the rate of degradation of these substances by MAO, although the activity of the aliphatic amino group is only slightly influenced by these substitutions (22a, b).

From these and many other data it has been concluded that in the active site of MAO an aromatic ring plus a two-membered side chain is located, which favors the attachment of phenethylamine or analogous compounds to the oxidase (23). When the eutopic enzyme-substrate complex is achieved, the removal of one α-hydrogen gets under way according to Eqs. ( 3 ) and ( 4 ) , but not after the formation of dystopic complexes.

3 . D I A M I N E O X I D A S E S

Although diamine oxidase (DAO) (23a) seems to occur more diversely in the whole biosphere than MAO, it is found in fewer mammalian organs than the latter oxidases. To name one instance, DAO is lacking in the brain of man (13) and other mammals (24). Within the cell it is usually localized in the soluble fraction; in rabbit liver, however, mitochondria contain most of the DAO activity (25, 26).

DAO, like the other amine oxidase, undergoes marked quantitative changes during ontogenesis (9). In addition, it appears to be closely con­

nected with reproduction. This is indicated by its presence in human prostatic gland, seminal plasma, placenta, and in blood plasma of preg­

nant women (9).

A number of related amine oxidases with partially overlapping sub­

strate patterns are known. For illustration we compare the classic hog kidney DAO and spermine oxidase. The former attacks cadaverine rapidly, spermidine much less, and spermine hardly at all (27). Exactly the reverse is true for spermine oxidase (28). All these amine oxidases share a remarkable sensitivity toward acylhydrazides, e.g., semicarbazide, isoniazid, and aminoguanidine. While p l5 0 values (negative logarithm of molarity producing 5 0 % inhibition) from 4 to 7 are found for these hydrazides, they are less than 1 or 2 for MAO. The situation is completely reversed for phenylcyclopropylamines (Section II, C) and for iV-ben- zyl-iV-methyl-2-propylamine that block MAO most efficiently ( p l5 0 val-

(6)

5 8 Ε. Α. ZELLER H2N ( C H2)6N H2

(Π) Cadaverine Η2Ν (CH2)3NH (CH2)4NH2

(HI) Spermidine

H2N(CH2)3NH(CH2)4NH(CH2)3NH2

(IV) Spermine

ues from 6 to 7) without affecting appreciably the semicarbazide-sensitive oxidases. Thus, in spite of many similarities, two distinct families of amine oxidases exist. It has been proposed to separate them operationally with the help of 1 0 ~3 Μ semicarbazide (23). To the semicarbazide- sensitive enzymes which are called diamine oxidases belong a number of catalysts which display somewhat greater differences among themselves than are seen in the group of monoamine oxidases. DAO in Mycobacte­

rium smegmatis (29), plant amine oxidases (30), porcine DAO [which was discovered as histaminase (31)], mescaline oxidase in rabbit liver

(26, 32), and spermine oxidase in blood plasma of ruminants (33) belong to this semicarbazide-sensitive family. Spermine oxidase was isolated recently in crystalline form and contained C u + + and pyridoxal phosphate (34), and pea seedling DAO was obtained in a rose red, highly purified form (35,85a).

Since the initial steps catalyzed by MAO and DAO are covered by the same equation [Eq. ( 1 ) ] , by analogy the interaction between diamines and DAO may follow a sequence similar to that outlined in Eqs. ( 3 - 5 ) . The binding of the substrate to the active site, however, is quite different from that found for MAO. Only one amine residue, called A, is released from the substrate by the enzymic process. It appears to be bound to an enzymic aldehyde residue (36, 37), presumably that of pyridoxal (receptor a). The second amino group, B, may be replaced by iV-alkylated residues, by imidazole (histamine), by other heterocyclic rings, and by guanidine (agmatine) (38). Either an organic electrophilic group or a metal possibly may act as receptor for group Β (receptor b). The distance between the receptors a and b largely determines whether a given diamine may serve as a substrate (39). Group Β is not essential for the substrate- enzyme interaction, since it can be left out. The affinity, as measured by the reciprocal of Km, however, drops sharply when the second nucleophilic residue is missing (40).

(7)

II. AMINE ANALOGUES

A. Polyamines a n d Mescaline

The polyamines spermine (IV) and spermidine (III) are widely distributed in biological materials, while diamines such as putrescine (1,4-diaminobutane) and cadaverine [1,5-diaminopentane ( I I ) ] have been positively identified in only a few instances. Since the two groups of amines often act in a similar fashion, albeit with different quantitative efficiency, the diamines can be considered in many cases as polyamine analogues. We still do not know the specific function of these amines, but considerable information has recently been accumulated on their biosyn- thesis and metabolism, as well as on a variety of physiological and pharmacological effects. Some of them serve as growth factors for certain microorganisms, and some as bactericidal substances. They produce vivid mating colors in certain fish and induce chromosomal aberrations in plants. In searching for the basis of their biological function one finds that polyamines interact with polyanions, such as nucleic acids and heparin, and that they possess the ability to influence membrane stability of bacteria, animals cells and cell particulates. Tabor, Tabor, and Rosenthal, have reviewed the field recently (4-0-

1. E N Z Y M E S P E C I F I C I T Y A N D BIOLOGICAL S T A B I L I T Y

Hog kidney, as first shown by Zeller, rapidly oxidizes putrescine and cadaverine. It slowly degrades spermidine and spermidine analogues, also by a semicarbazide-sensitive process of oxidative deamination (27).

Spermine, however, is not attacked by purified D A O ; crude kidney preparations from pork, guinea pig (42), and sheep (43) display a minute capacity for deaminating spermine. In pork kidney extracts the oxida- tive deamination starts only after a long lag period (36, 44)- A strong spermine-destroying power is observed in many bacteria, as first found in Pseudomonas pyocyanea by Silverman and Evans (45), and in M. smeg- matis by Roulet and Zeller (29). In the latter case the semicarbazide sensitivity of the enzymic process was ascertained. Outside the micro- biological realm the only well-defined spermine-attacking enzyme is the spermine oxidase occurring in the blood plasma of ruminants (45a) (Section I, B, 3 ) . In searching for the origin of this plasma enzyme, Barsky could not find any appreciable spermine oxidase in the homoge- nates of six organs from sheep and cattle (and four other species) (43).

Why are MAO and DAO unable to deal with polyamines although

(8)

60 Ε . Α. ZELLER

these compounds possess the basic structure of formula (I)? The same question can be asked about mescaline which, as a derivative of phenyl- ethylamine (3,4,5-trimethoxyphenylethylamine), seems to be ideally suited to serve as a substrate for MAO. Since mescaline is susceptible to DAO (26) and spermine to spermine oxidase, the inability of MAO to catalyze the removal of certain amino residues cannot be accounted for by an abnormally high stability of the carbon-nitrogen bond. Moreover, the nonsubstrate amines compete efficiently with the ordinary substrates, indicating that the former are actually bound to the active site of MAO (22a, b). Obviously, the suitable amines form eutopic complexes which lead to the decomposition of the substrate, whereas the resistant sub­

stances enter into a dystopic complex which does not break down to the enzyme and reaction products. A model of the dystopic binding of short- chained diamines by MAO has been presented before (46). According to Zeller and his group, even good substrates of a given enzyme may produce in part dystopic complexes. The relative degree of dystopic complex formation may be evaluated with the help of a modification of the Michaelis-Menten relationship [Eq. ( 6 ) ] , where the subscripts d and e

refer to dystopic and eutopic complexes and l / δ is defined by (Kd + Ke)/Kd. If Ke = Km, meaning no appreciable amount of dystopic com­

plexes is present, Eq. (6) reverts to the standard form of the Michaelis- Menten relationship. If the dystopic complexes prevail (Ke>> Kd), then ν approaches zero, as in the system mescaline/ΜΑΟ (22a, b).

The enzymic resistance of spermine toward MAO may well be one factor accounting for the amine's ubiquity in animal tissues. Similarly, mescaline, when taken orally (as a part of the cactus Anhalonium), would never get past the active MAO of the intestinal mucosa, liver, brain capil­

lary, and brain cell, and never produce the classic hallucinations if it were not dystopically attached to this enzyme.

2. I N T E R A C T I O N B E T W E E N P O L Y A M I N E S A N D E N Z Y M E S

D i - and polyamines form complexes with the polyanionic nucleic acids, as evidenced by the pronounced effect of aliphatic diamines on the transi­

tion of the helix of nucleic acids to the random coil (47) and by the protective effect of spermine and other polyamines against heat denatura- tion of D N A (48). It seems possible that the production of these com­

plexes is responsible for the inhibitory power of spermine on the hydrolysis of R N A and D N A catalyzed by purified nucleases (19).

(9)

Activation of a bovine testicular hyaluronidase by spermine, spermi­

dine, and, to a lesser extent, by several other amines has been demon­

strated by Miyaki et al. (50). In this study, oxidation of spermine by goat serum amine oxidase destroyed its activating ability. In view of the previous reports that hyaluronidase may play a role in the penetration of the egg cell by spermatozoa, these authors pointed out that the activa­

tion of hyaluronidase by spermine may have significance for fertilization.

When sheep plasma is incubated with spermine, a strong denaturation of plasma proteins ensues ; (48). This simple observation displays the strong biological effects of the products of the system spermine/spermine oxidase. Many more complex phenomena may have a similar basis. With Mycobacterium tuberculosis, for example, Hirsch and Dubos showed that the toxic agent was not spermine but rather the oxidation product of spermine that resulted from the action of beef serum amine oxidase

(83, 42).

B y no means is this kind of data restricted to the field of diamines and polyamines. To mention two examples, some products of tyramine/MAO inhibit choline oxidase (51) and succinodehydrase (52). Quastel pointed out a long time ago that products of enzymic action rather than the amines themselves may be responsible for at least some of the biological effects observed after the administration of amines.

3. I N C O R P O R A T I O N OF A M I N E S I N T O P R O T E I N S

In Section I, Β, 1 the incorporation of amines into proteins with the help of transglutaminase was discussed. The task before us remains to indicate which amines have been found to be partners in this re­

action. Putrescine and cadaverine are excellent substrates and, to a lesser degree, spermine and mescaline (7). Whether amines are bound and stored this way, whether the incorporation of amines changes perceptibly the biological activity of specific proteins, or whether synthetic com­

pounds can be built into cell proteins in this fashion remains to be seen.

Pharmacological endeavor may one day open a rich field here.

B. a-Alkylamines

Amines alkylated in the α-position are not substrates of MAO or DAO (9,10). This often-confirmed behavior of compounds such as amphetamine (V) and ephedrine (VI) is considered to be in part responsible for the long duration of their pharmacological action. As indicated by competi­

tion experiments, the α-alkylamines are still capable of attaching them-

(10)

62 Ε. Α. ZELLER

0 ~

c h CH

*

2—CH—ΝΗCH3 2 CH—CH—NHCH, OH CHS

(V) (VI)

selves reversibly to the active site of MAO. Their in vitro inhibitory power, however, is rather small. Under standard conditions it takes 1 0- 2- 1 0 ~3 Μ concentrations to cut down the MAO activity to half of its original value, as compared with 1 0- 5- 1 0 ~7 Μ concentrations for modern inhibitors, such as iproniazid and phenylcyclopropylamines. The term inhibitor was taken much too literally and without regard to quantitative relationships by many nonenzymologists. In one instance, a one to a million dilution of ephedrine was used, which was considered to be suffi­

cient to eliminate MAO in an isolated aortic strip. An angry glance at the enzyme, however, would have been as effective as this dilute ephedrine solution. Only a few in vivo inhibition experiments with a-alkylamines were carried out, and they turned out to be negative (53). With the coming of isotopically labeled compounds, more refined tests on the in vivo action of ephedrine and related sympathomimetic drugs could be designed. Schayer was the first to show that the metabolism of several radioactive amines, including epinephrine, was not greatly influenced by ephedrine and amphetamines, but quite substantially by ipronia­

zid (54, 55). Since labeled amines are of necessity "exogenous" and since endogenous amines often suffer a metabolic fate different from that of exogenous substances, these experiments do not entirely preclude an effect of these α-alkyl derivatives on the metabolism of endogenous amines.

This reservation was overcome by pretreating rats with amphetamine, ephedrine, or iproniazid, and by subsequent determination of their brain serotonin; of the three, only iproniazid raised the serotonin level of the brain, whereas the first two were ineffective in this respect (56). Further­

more, the central nervous system appears to react perceptibly only when more than half of the brain MAO activity is eliminated. While in vivo permeability and distribution factors may play a role, it still is very difficult to visualize ephedrine reducing brain MAO by more than 5 0 % .

The phenylcyclopropylamines represent an impressive example of the fact that compounds closely related to amphetamine may display tre­

mendous inhibitory power for MAO (Section II, C). A promising step in bridging the enormous gap between the blocking efficiencies of ampheta­

mine and the phenylcyclopropylamines was made when advantage was taken of the high affinity of indole-3-ethylamine (tryptamine) for the oxidase (57) and when α-methyltryptamine and a-ethyltryptamine

(11)

(etryptamine) were synthesized and tested. The latter compound was found to be a stronger MAO inhibitor than amphetamine ( p l5 0 ~ 4) (58).

At long last, a substantial in vivo inactivation of MAO (rat liver) with an α-alkylamine was obtained (58). Moreover, after treating animals with etryptamine and 5-hydroxytryptophan, Greig et al. noticed a marked increase in the brain levels of serotonin and epinephrine (58). Thus, there seems to be little doubt, at least in the case of the rat, that a-ethyltrypta- mine, by causing accumulation of certain amines to a high degree, can produce pharmacological effects. Etryptamine actually increases motor activity, as measured by the actophotometer, and prevents reserpine- induced depression (58).

In rabbits, immediately following the administration of a-methyltrypta- mine, the electroencephalogram (EEG) tracing resembled that of the arousal EEG produced by amphetamine. While the initial changes in the EEG occurred long before there was an accumulation of serotonin, delayed action of the amine on the EEG seemed to correlate with the increase and decrease of serotonin (59).

C. Phenylcyclopropylamines (PCP)

Burger et al. (60), in synthesizing phenylcyclopropylamines (VII), expected to produce substances with amphetamine-like activity. The pharmacodynamic properties of the cyclopropylamine derivatives, how­

ever, were quite distinct from that of amphetamine and were rather sug­

gestive of MAO inhibition (61); this assumption was soon supported by the results of in vitro experiments (62, 63). Since then a large number of enzymic, metabolic, pharmacological, and clinical data concerning the great inhibitory power of phenylcyclopropylamines has been reported

(68a).

1. In vitro E X P E R I M E N T S

Cis- and trans-PCP belong to the most effective in vitro inhibitors of MAO (62, 63), as indicated by the p l5 0 values, which are the highest ever recorded (20). From the data of Table I, one has to conclude that the

C H2

(VII)

(12)

64 Ε . Α. ZELLER

absolute and relative blocking powers of the two geometric isomers varies with the origin of the MAO preparation. Although DAO in many ways is closely related to MAO, the former enzyme is not affected by cis- and trans-PCP (Section I, B, 3). Furthermore, another enzyme catalyzing

T A B L E I

SPECIES AND ORGAN SPECIFICITY OF M A O INHIBITION BY PHENYLCYCLOPROPYLAMINES

(pl

5 0 VALUES)

Origin of MAO CVs-PCP Trans-FCF

Beef liver 6.8 7.0

Rabbit liver 6.0 5.7

Rabbit brain

6.3

oxidative deamination, L-amino acid oxidase of snake venoms {20), sev­

eral other dehydrogenases {64), and various other enzymes {20) turned out to be insensitive toward PCP.

While PCP and hydrazine derivatives, e.g. iproniazid, share many inhibitory properties, fundamental differences were discovered on closer inspection {19). The data are compatible with the assumption that iproniazid is bound to MAO by a covalent bond [see Section II, D , 1 and formula ( I X ) ] , whereas for the binding of PCP, only secondary forces come into play {19,21).

2. In vivo E X P E R I M E N T S

In rabbits the in vivo inhibition of MAO by P C P can be observed easily. The activity of the brain enzyme is reduced more strongly and for a longer period than that of liver MAO (20), a fact which may be related to the high in vitro sensitivity of brain MAO (Table I ) . It takes 5 days for the rat brain to recover its original enzymic activity after the admin­

istration of trans-PCP {24). Apparently, the in vivo and in vitro in­

hibitory specificities are much the same, since 50 times the quantity of trans-PCP which produces 50% inactivation of MAO does not alter DAO activity in cat kidney {24).

Although the metabolic effects of MAO inhibition have not been thor­

oughly explored as yet, some decisive observations have been reported.

(a) In rat brain, the level of norepinephrine increases by 80-90% 5 hours after the administration of a small amount of trans-PCP {65). (b) Green et al. find brain serotonin to be consistently increased after the applica­

tion of this drug {66, 67). (c) The urinary excretion of tyramine and

(13)

epinephrine in dog (68) and of tryptamine in man (69) is enhanced by tranylcypromine (systematic name for irans-PCP), whereas that of 5-hydroxyindoleacetic acid is decreased (70). Since MAO appears to be mostly responsible for the degradation of tryptamine, we are not surprised to learn that trans-PCP potentiates the convulsion-producing power of this amine (71).

In a number of cases, a close correlation between PCP-induced meta- bolic changes and biological effects have been noted. Two examples are given: (a) The ability of P C P to arouse rats from reserpine-induced sedation was closely correlated with the restoration of the brain's capacity to accumulate serotonin without a concomitant increase of the norepine- phrine level (67). (b) A rapid rise in the serotonin level of rabbits after PCP administration seems to be related to the desynchronization of the EEG (68).

It has to be emphasized that MAO is not necessarily the only target of the phenylcyclopropylamines, even if no other defined receptor is known as yet. One well-analyzed case will demonstrate this point. The positive inotropic action of trans-PCP (and other MAO inhibitors) appears pri- marily to be due to its capacity to release endogenous catechol amines from their cardiac stores (72). The duration of the response is five times longer in the rat than in the cat heart. This species difference may be explained by the fact that, in homogenates of the ventricular myocardium of the rat (and man), the catabolism of catechol amines is strongly blocked by PCP, but no such influence is seen in cat preparations. In the hearts of rat and man, therefore, irans-PCP seems to have a twofold function. It releases catechol amines and protects these compounds against destruction by MAO, whereas in cat heart only the former action comes into play (see also Section II, D , 4c).

D. Hydrazine Derivatives

When, in 1952, the hydrazine derivative iproniazid, 1-isonicotinic acid 2-isopropylhydrazide (VIII) was found to be an efficient inhibitor of

MAO in vitro (73) and in vivo (74), a new era in the study of the enzymology, metabolism, general pharmacology, and psyehopharmacology of amine metabolism was ushered in. In 1960 Pletscher, Gey, and Zeller

(vin)

(14)

66 Ε. Α. ZELLER

published a review of this rapidly growing field, based on more than 1300 references (75); the flood of investigations still does not show any sign of recession. While the period from 1952 to 1960 is thoroughly treated in Pletscher's review, unfortunately the time since 1960 can be covered here only by reference to a handful of papers, which were chosen for the purpose of demonstrating some applications of the new tool to the study of a wide variety of biological problems. Additional data are found in a more recent review (63a).

Hydrazines had to be considered as a family of purely synthetic compounds until recently, when Levenberg isolated agaritine, a phenyl- hydrazide of L-glutamic acid (β-Ν- [ y - L ( - f ) -glutamyl ]-4-hy droxy- methylphenylhydrazine) from Agaricus bisporus, the most common edible mushroom (76). In the past, the group of hydrazine derivatives did not receive much pharmacological attention. Phenylhydrazines, however, have been used for some time to produce anemia (by an unknown mecha­

nism), phenylalkylhydrazines to exert sympathomimetic activity similar to that of the corresponding phenylalkylamines (77), and hydrazino phthalazines to reduce blood pressure (78).

Hydrazines not only display the same chemical reactivity as amines (Section I,A) but also participate in many oxidoreductive reactions.

Above all, they furnish the bulk of carbonyl reagents. They share this property with hydrazine itself and with many acylhydrazides. But neither these compounds nor potassium cyanide influence the primary step in the MAO reaction, indicating that no easily accessible carbonyl residue is present in this enzyme and that hydrazine derivatives owe their tremen­

dous MAO-blocking effect not to their ability to react with ketones and aldehydes, but to other physicochemical properties to be mentioned in the subsequent discussion (79).

1. In vitro S T U D I E S

How do hydrazines influence MAO if not as carbonyl reagents? The answer to this question came from an investigation of the action of iproniazid. Zeller et al. observed that substrates compete with iproniazid (82) and that the — N H N H C H ( C H3)2 moiety of iproniazid seems to be responsible for the action of the latter (80, 81). Furthermore, if the isosteric replacement of C H2 by N H residues is kept in mind, similarities between the substrate and inhibitor patterns unmistakably appear (83, 84). Since the degree of inhibition is a function of the amount of time of preincubation of the enzyme preparation with iproniazid (before the sub­

strate is added), and since isopropylhydrazine and other alkylhydrazines are much more effective than iproniazid, the hydrolysis of iproniazid to

(15)

form isopropylhydrazine and isonicotinic acid prior to the inhibitory action has been suspected. This concept is supported by some recent observations. Labeled isonicotinic acid was isolated after the incubation of iproniazid with solubilized rat brain mitochondria (85), and labeled benzylhydrazine was found in rats treated with isocarboxazid (86).

These are only a few of the many available data which lead us to believe that the hydrazine inhibitors (or some metabolic degradation products thereof) occupy the same part on the active site of MAO as the corre­

sponding substrates. Once the hydrazine is attached to the active site by secondary forces, a covalent bond is perhaps established. Three sets of data favor this idea: (a) Davison's important observation that no inhibi­

tion of MAO by iproniazid is produced in the absence of oxygen (19, 87);

(b) the inability of iV'-disubstituted hydrazines to block the oxidase (80, 81, 88); and (c) the lack of success of every attempt to reverse the iproniazid-included inhibition (19, 82). All these results are consistent with the assumption that the hydrazine derivatives, after being attached to the active site by secondary forces in the manner of substrates

(Eq. (3), Section I, B, 2 ) , form a covalent bond with the enzyme ( I X ) . The reaction sequence, however, does not proceed beyond this point, and the hydrazine remains attached to the ΜΑΟ ( I X ) . In three recent papers

R — N N H2

I

E N (IX)

the blocking pattern of hydrazine inhibitors has been thoroughly studied, and much new light has been thrown on the structure-activity relation­

ships (89-91).

There have been many occasions when MAO inhibitors were used for enzymological studies. For example, effective inhibitors are useful when it is necessary to establish whether an unusual amine, which is oxidized by the enzyme preparation, is a substrate of MAO or of an accompanying enzyme (63a). With the help of iproniazid and a number of nonhydrazine agents it was proven that the oxide of iV-dimethyltryptamine (92) and γ-aminobutyronitrile are really attacked by MAO.

2. In vivo E X P E R I M E N T S

The term "in vivo inhibition" is accepted here only when the drop of MAO activity after the administration of the inhibitor has actually been determined. A rise of the serotonin level in brain may be a consequence of the MAO block, but does not per se constitute an unambiguous proof

(16)

68 Ε . Α. ZELLER

of the in vivo inhibition. Reactions so defined were first demonstrated by Zeller and Barsky (74)· In addition, these authors, together with Berman, followed the recovery of MAO activity in rats until, after 96 hours, the original enzyme level was reached (82). This prolonged action cannot be due to the survival of iproniazid, since this compound is rapidly destroyed and disappears at a time when the level of MAO activity is at its lowest (93).

Another facet of in vivo inhibition was studied by Pletscher (75, p. 461) and by McGrath and Horita (94). These authors found that the relative degree of the MAO block in brain and liver depends considerably on the structure of the agents used and on the way they have been administered;

iproniazid preferentially affects liver MAO, while phenylisopropylamine (pheniprazine) reduces brain MAO more than liver enzyme activity.

Since iproniazid appears to block MAO irreversibly (Section II, D , 1) the restoration of the enzymic activity in the course of several days may be a feedback-controlled synthesis rather than a reactivation of the inactivated enzyme molecule. This concept is supported, but not proven, by recent results obtained by Zeller, Lane, and Schweppe, who found that the rate of recovery of the MAO activity in liver after a single iproniazid injection into male mice is significantly enhanced by castration. For more data the reader is reminded of the comprehensive review given in refer­

ence 75 (p. 462).

3. MAO I N H I B I T O R S AS T O O L S I N BIOLOGICAL R E S E A R C H

Amine analogues in the form of hydrazine derivatives have been used to study the biology of many organisms, from protozoa to spiders to man.

The following examples can give only a rather limited view of the diversity of problems to which MAO inhibitors have been applied.

a. Lathyrism. Lathyrism, a nutritional collagen disease, can be pro­

duced experimentally with γ-aminopropionitrile. In rats the severity of the symptoms of experimental lathyrism is aggravated by iproniazid.

Since γ-aminopropionitrile is deaminated by MAO, albeit slowly (Section II,D,1), this effect of iproniazid could be due to protection of the nitrile from enzymic destruction. Further investigation, however, revealed that isoniazid was also effective in potentiating the lathyrogenic effect of γ-aminopropionitrile (95, 96). This result all but excludes MAO from further consideration because isoniazid, differing from iproniazid only by the absence of the isopropyl residue, does not inactivate MAO. It was fortunate that iproniazid and isoniazid became available simultaneously and that the differences in MAO-blocking power was soon discovered

(17)

(78, 74, 81, 97, 98). Since that time they frequently have been used to decide whether MAO was involved in producing certain specific phe- nomena (Section II, D , 4c).

b. Protection against X-Rays. Since tryptamine and 5-hydroxytrypta- mine (serotonin) are "powerful radioactive amines," it should be possible to enhance their protective action by blocking MAO. A reaction of this kind may have occurred when van den Brenk et al. treated rats with iproniazid and reserpine (99). These authors observed that reserpine, if given immediately before irradiation, caused a slight protection, which could be increased by pretreating the animals with iproniazid. It would be of interest to know the results of treatment of experimental animals with a combination of MAO inhibitors and one of the indoleethylamines.

c. Metabolism of Exogenous Dopamine. Dopamine (3,4-dihydroxy- phenylethylamine, 3-hydroxytyramine) is degraded through various channels. The selection of a given metabolic sequence may depend on the method of administration of the biogenic amine or on the individual tissue in which the transformation takes place. One may obtain a gen- eral view of dopamine metabolism by determining the radioactive urinary excretion products derived from labeled dopamine. This was done by Williams et al. (100), who repeated and extended the work of Goldstein et al. (101). The authors injected dopamine intraperitoneally into rats and determined the excretion of labeled (a) homoprotocatechuic acid,

(b) dopamine, (c) methoxytyramine, and (d) homovanillic acid. One may assume that homoprotocatechuic acid, by way of the corresponding aldehyde, is a product of MAO action, that methoxytyramine is a result of catechol O-methyltransferase activity, and that homovanillic acid is formed with the help of both enzymes. This concept was borne out by experiments in which the animals were treated with isocarboxazid

[ l-benzyl-2- (5-methyl-3-isoxazolylcarbonyl) hydrazine]. The inactiva- tion of MAO by this drug led to a threefold increase in the urinary excretion of methoxytyramine and to a tenfold reduction of the excretion of homovanillic and homoprotocatechuic acids. From these and additional data it was "concluded that both O-methylation and oxidative deamina- tion are of equal importance for the metabolism of exogenous dopamine."

Since pyrogallol, an inhibitor of the O-methylating enzyme in doses which almost completely blocked the formation of 3-methoxytyramine, did not prevent the formation of homovanillic acid, one has to assume that O-methylation of exogenous dopamine is not an obligatory step before the onset of oxidative deamination.

d. Metabolism of Endogenous Serotonin in the Central Nervous Sys- tem. Serotonin, when it is attacked by MAO, is converted into indoleacet-

(18)

70 Ε. Α. ZELLER aldehyde, which in turn is oxidized by aldehyde dehydrogenases into 5-hydroxyindoleacetic acid (5-HIAA). Recently, the distribution of 5-HIAA in various parts of the rabbit brain was determined by Roos (102). Twenty hours after the administration of the MAO inhibitor nialamide (1- [2- (benzylcarbamyl) ethyl] -2-isonicotinoylhydrazine), the 5-HIAA completely disappeared. Obviously, endogenous serotonin in the rabbit brain is deaminated by MAO.

e. Endogenous versus Exogenous Norepinephrine. A large number of investigations have shown that norepinephrine is metabolized largely by two alternate enzymes: catechol O-methyltransferase and MAO. Crout et al. undertook a penetrating study to evaluate which of these two en­

zymes is of greater significance in the metabolism of norepinephrine in the brain and heart of the rat (103). Three types of experiments were per­

formed. Initially, the enzyme activities of both methyltransferase and MAO in the brain, heart, and liver were assayed in vitro to evaluate the capacity of each tissue to carry out each of these metabolic steps. MAO activity was found to be greater than that of the catechol O-methyl- transferase in the brain and in the heart, whereas the reverse was true in the liver. Secondly, inhibitors of the transferase and of MAO were administered in vivo, separately and together, to determine which caused a greater accumulation of endogenously formed catechol amines in the brain and in the heart. A significant increase in the catechol amine content of rat brain and heart was produced by the intraperitoneal administration of four different MAO inhibitors, whereas the injection of pyrogallol, an inhibitor of the transferase, failed to raise the level of catechol amines in these organs. Finally, the in vivo accumulation of injected norepine­

phrine by rat heart was studied in control animals and in animals treated with an inhibitor of each enzyme to determine the relative ability of each enzyme to degrade exogenous norepinephrine that penetrates into the myocardium. In animals pretreated with iproniazid the major finding was a marked accumulation of amine in the myocardium, relative to the con­

centration in plasma. Rats pretreated with pyrogallol, on the other hand, demonstrated a severe impairment in their ability to metabolize circu­

lating norepinephrine. The results of these experiments suggest that MAO plays the greater role in the initial metabolism of norepinephrine in the brain and heart, while the catechol-O-methyltransferase is of greater significance in the liver. Since circulating norepinephrine is carried by the blood to the two organs with the highest concentration of the methyl­

transferase, liver and kidney, the importance of this enzyme for the destruction of the circulating and (labeled) exogenous norepinephrine becomes apparent.

(19)

/. Analysis of Drug Action: Rauwolfia Alkaloids and Chlorpromazine.

Soon after Rauwolfia alkaloids were known to deplete brain and other organs of serotonin and catechol amines, the research groups of Brodie and of Pletscher began to study these effects with the help of MAO inhibitors. It was found that iproniazid not only counteracts the reserpine- induced sedation, but also prevents the loss of biogenic amines from various tissues (75). Since that time the pharmacological studies on Rauwolfia alkaloids and on MAO inhibitors have become intimately interwoven and the antireserpine effects have been used for the large scale in vivo screening of potential MAO inhibitors. The paper by Bickel et al.

(104) is an example of a recent study of this type. These authors showed that the single injection of two chemically fairly different inhibitors, nialamide and I.S. 2596 [iV'-(l,4-benzodioxane-2-methyl)-N'-benzyl- hydrazine] prevented the reserpine-induced increase in the urinary catechol amines. This action of the two inhibitors lasted for 1 week, the same duration as their blocking effect on MAO.

Inhibitors of MAO were instrumental in the discovery of a new and important facet of the mechanism of action of chlorpromazine. Camanni et al. (105), Ehringer et al. (106), and Pletscher et al. (107) observed that high doses of chlorpromazine prevented the protection afforded by iproniazid against the reserpine-caused release of catechol amines in the adrenal medulla and in the brain. A thorough analysis of this phe­

nomenon was recently published by Gey et al. (108). Chlorpromazine alone did not affect the levels of serotonin and norepinephrine in the rat brain, but it counteracted the iproniazid-produced increase of the con­

centrations of these two amines. The authors suggest that chlorpromazine may decrease the permeability of the storage organelles for these amines.

In this field of investigation, the usefulness of the MAO inhibitors stems from their ability to raise the amine level far above the one caused by the Rauwolfia alkaloids. Within the wide range of these two levels the chlor­

promazine effect becomes easily discernible.

g. Discovery of Ο dopamine (Norsympathol) in the Mammalian Or­

ganism. In view of the ubiquity of MAO, it seems quite feasible that more than one biogenic monoamine hitherto has escaped detection. Actually, the previously doubtful existence of tyramine and tryptamine in mam­

mals was finally confirmed with the help of MAO inhibitors by Sjoerdsma et al. (109). Recently, p-hydroxymandelic acid was discovered in the urine of several mammalian species, including man. Kakimoto and Arm­

strong suspected that this acid may have been produced from octopamine (X) by the action of MAO (110). After treating rabbits with iproniazid or phenylisopropylamine, the authors not only observed the appearance

(20)

7 2 Ε . Α. ZELLER

of octopamine in the urine, but also found this compound to become the most prominent phenolic amine in the extracts of several organs. In

the future, when the biological effects of MAO inhibitors are evaluated, the proportionately large changes in the metabolism of phenolic amines have to be kept in mind.

4 . C O M P L E X I T Y OF B I O C H E M I C A L P H E N O M E N A

Amines and amine analogues are capable of interacting with so many components of living organisms (Sections I, A and II, D ) that one has to be cautious in the interpretation of phenomena induced by amines.

Even if one limits the discussion to substances acting only on one target enzyme, a great number of parameters have to be considered. One group of factors, summarized under the term accessibility, is comprised of the pharmacon's absorption, transport, metabolic destruction, the intracellu­

lar concentration gradients, and organ and species specificity. Other com­

plicating factors are discussed in the following sections.

a. Competition. In various biological situations probably more than one biogenic amine is inactivated by amine oxidases. Therefore, we have to expect different degrees of inhibition, depending on the substrate which at the given moment is being attacked by the enzyme. The wide range of these quantitative differences may be inferred from the pair tyramine/

MAO and phenylbutylamine/MAO; the former is blocked in vitro 2 0 0 - 4 0 0 times more strongly by phenylcyclopropylamines than the latter (19).

There is such a close structural and isosteric relationship between iproniazid and epinephrine that another type of competition, this time for the receptor site of the effector cell, might be considered. This concept was helpful in the evaluation of the experiments of Griesemer et al. (Ill) and of Kamijo et al. (112). The former produced a reversible adrenergic block in the isolated rabbit aorta with the help of iproniazid, while the latter studied the effect of this drug on the response of the cat nictitating membrane to the stimulation of the sympathetic trunk.

b. Metabolic Conversion of Inert Compounds into Active Inhibitors.

In view of Bovet's observation that phenylethylhydrazines display sym­

pathomimetic activity (Section II, D , and ref. 77), it is not surprising to find direct, amphetamine-like stimulation of the CNS by several MAO

(X)

(21)

III. CONCLUDING REAAARKS

This short review, in spite of its obvious limitations, should demon- strate how manifold the interaction among amine and amine analogues appear to be. Biogenic amines compete with each other for chemorecep- inhibitors of similar chemical structure, e.g., phenylisopropylhydrazine

(pheniprazine) (113) and a-ethyltryptamine (114)- The fact that ipronia- zid is devoid of analeptic power supports the previously mentioned as- sumption that iproniazid is converted to isopropylhydrazine prior to its in vivo action. The adrenolytic effect mentioned above (Section II, D , 4a) is more likely to result from the intact iproniazid molecule.

c. Membrane Effects. Since 1957 it has been suspected that iproniazid and other MAO inhibitors act on amine metabolism not only by blocking MAO, but also by preventing endogenous amines from leaving the cell.

Since that time this hypothesis has been used in the interpretation of a number of experimental data (115). Axelrod et al. (116), on finding that three MAO inhibitors of different chemical structure and duration of action (phenylisopropylhydrazine, iV-methyl-iV-benzyl-2-propynylamine, and harmaline) blocked the releasing action of reserpine on norepine- phrine-H3 in the rat heart, proposed that the MAO inhibitors elevated the catecholamines in certain tissues by blocking the release of norepinephrine from its binding site. Pepeu et al. compared the action of iproniazid and phenylisopropylhydrazine (117). Whereas both compounds completely destroyed MAO of guinea pig heart atria, only iproniazid was able to counteract some of the depressant effect of reserpine. These authors ex- plained their observations by assuming that only iproniazid presented a resistance to the release of amines in addition to being a MAO inhibitor.

Spector et al., however, did not believe that MAO inhibitors per se pre- vent the release of biogenic amines from cells (118). They based their conclusion on three observations: (a) MAO inhibitors do not prevent the release of serotonin from rabbit platelets in vivo or in vitro; (b) tem- porary blocking with harmaline, a short-acting competitive MAO inhibi- tor, does not block the release of brain amines by reserpine; (c) a variety of MAO inhibitors of various structures preserve the pharmacological action of reserpine and prevent the decline in amine level. Furthermore, whenever isoniazid was tested (103, 105, 111, 112), it did not influence the disappearance of amines from cells. Apparently, it has not been definitely determined whether inhibitors resist the release of amines by interacting with MAO.

(22)

7 4 Ε. Α. ZELLER tors and for the binding sites of certain enzymes, and thus tend to poten­

tiate or reduce each others' biological activities. More complications are introduced by the power of certain amines to release other compounds, e.g. catechol amines, with ensuing interplay between releasing and re­

leased amines, or by competitive events between endogenous and exog­

enous amines. When we add hydrazine derivatives to our list, we find some of these isosteric amines competing with ordinary amines for receptor sites. In passing, the tremendous inhibitory effect of many hydrazines on amine oxidases and amino acid decarboxylases, and thus on amine formation and degradation, are mentioned again. With cyclo- propylamines, not only the degree but also the type of inhibition is changed profoundly when one substrate of monoamine oxidase is replaced by a closely related one. We are, therefore, not too surprised when we find that only part of the metabolic and functional network existing between amine and amine analogues is unraveled. On the other hand, the new tools, such as the monoamine oxidase inhibitors, have yielded remarkable experimental results and new insights into the mode of action of amine analogues which in turn stimulated the undertaking of many valuable clinical studies.

REFERENCES

1. H. G. Grimm, M. Gunther, and H. Titus, Z. physik. Chem. Abt. B14, 169 (1931).

2. H. Erlenmeyer, Bull. soc. chim. biol. 30, 792 (1948).

3. R. Ito, T. Ito, and K. Miyana, Nippon Univ. J. Med. 2, 459 (1960).

4. V. Fishman and H. Goldenberg, Proc. Soc. Exptl. Biol. Med. 110, 187 (1962).

5. W. Block, Z. physiol. Chem. Hoppe-Seyler's 296, 108 (1954).

6. Ν. K. Sarkar, D. D. Clarke, and H. Waelsch, Biochim. et Biophys. Acta 25, 451 (1957).

7. H. Waelsch, in "Chemical Pathology of the Nervous System" (J. Folch-Pi, ed.), p. 576. Pergamon Press, N e w York, 1961.

8. M. J. Mycek, D. D. Clarke, A. Neidle, and H. Waelsch, Arch. Biochem.

Biophys. 84, 528 (1959).

9. E. A. Zeller, in "The Enzymes" (J. B. Sumner and K. Myrback, eds.), Vol. II, p. 536. Academic Press, N e w York, 1951.

10. H. Blaschko, Pharmacol. Revs. 4, 415 (1952).

11. A. N. Davison, Physiol. Revs. 38, 729 (1958).

11a. C. Ε. M. Pugh and J. H. Quastel, Biochem. J. 31, 2306 (1937).

12. C. de Duve, H. Beaufay, P. Jacques, Y. Rahmah-Li, Ο. Z. Sellinger, R.

Wattiaux, and F. de Coninck, Biochim. et Biophys. Acta 40, 186 (1960).

13. H. Birkhauser, Helv. Chim. Acta 23, 1071 (1940).

14. D. F. Bogdanski, H. Weissbach, and S. Udenfriend, J. Neurochem. 1, 272 (1957).

(23)

15. M. Eder, Beitr. pathol. Anat. u. allgem. Pathol. 117, 343 (1957).

15a. E . A. Zeller, R. E. Lane, and J. S. Schweppe, Endocrinology (to be pub­

lished) (1963).

16. W. J. Novick, Jr., Endocrinology 69, 55 (1961).

17. P. Hagen and N. Weiner, Federation Proc. 18, 1005 (1959).

18. S. Sarkar and E. A. Zeller, Federation Proc. 20, 238 (1961).

19. E. A. Zeller and S. Sarkar, J. Biol. Chem. 237, 2333 (1962).

20. S. Sarkar, R. Banerjee, M. S. Ise, and E. A. Zeller, Helv. Chim. Acta 43, 439 (1960).

21. B. Belleau, J. Burba, M. Pindell, and J. Reiffenstein, Science 133, 102 (1961).

22. G. A. Alles and Ε. V. Heegaard, J. Biol. Chem. 147, 487 (1943).

22a. E . A. Zeller, Ann. N.Y. Acad. Sci. 107, 811 (1963).

22b. E. A. Zeller, Biochem. Z. (to be published) (1963).

23. E. A. Zeller, Pharmacol. Revs. 11, 387 (1959).

23a. E. A. Zeller, in "The Enzymes" (P. D. Boyer, H. A. Lardy, and K. Myr- back, eds.), Vol. VIII, p. 313. Academic Press, N e w York, 1963.

24. W. P. Burkard, K. F. Gey, and A. Pletscher, Biochem. Pharmacol. 11, 177 (1962).

25. G. C. Cotzias and V. P. Dole, J. Biol. Chem. 196, 235 (1952).

26. E. A. Zeller, J. Barsky, E. R. Berman, M. S. Cherkas, and J. R. Fouts, J. Pharmacol. Exptl. Therap. 124, 282 (1958).

27. E. A. Zeller, B. Schar, and S. Staehlin, Helv. Chim. Acta 22, 837 (1939).

28. C. W. Tabor, H. Tabor, and S. M. Rosenthal, J. Biol. Chem. 208, 645 (1954).

29. F. Roulet and E. A. Zeller, Helv. Chim. Acta 28,1326 (1945).

30. R. H. Kenten and P. J. G. Mann, Biochem. J. 50, 360 (1962).

31. C. H. Best and E . W. McHenry, J. Physiol. (London) 70, 349 (1930).

32. F. Bernheim and M. L. C. Bernheim, J. Biol. Chem. 123, 317 (1938).

33. J. G. Hirsch and R. J. Dubos, J. Exptl. Med. 95, 191 (1952).

34. H. Yamada and Κ. T. Yasunobu, J. Biol. Chem. 237, 1511 (1962).

35. E. Werle, I. Trautschold, and D. Aures, Z. physiol. Chem. Hoppe-Seyler's 326, 200 (1961).

35a. P. J. G. Mann, Biochem. J. 78, 623 (1961).

36. E . A. Zeller, Helv. Chim. Acta 21, 1645 (1938).

37. Ε. V. Goryachenkova, Biochemistry (U.S.S.R.) (English Transl.) 21, 249 (1956).

38. E. A. Zeller, Advances in Enzymol. 2, 93 (1942).

39. E. A. Zeller, J. R. Fouts, J. A. Carbon, J. C. Lazanas, and W. Voegtli, Helv. Chim. Acta 39, 1632 (1956).

40. J. R. Fouts, L. A. Blanksma, J. A. Carbon, and E. A. Zeller, J. Biol.

Chem. 225, 1025 (1957).

41. H. Tabor, C. W. Tabor, and S. M. Rosenthal, Ann. Rev. Biochem. 30, 579 (1961).

42. J. G. Hirsch, J. Exptl. Med. 97, 327 (1953).

43. J. Barsky, On the Reactive Site of Monoamine Oxidase, Doctoral Thesis, Northwestern University, Evanston, Illinois, 1958; Dissertation Abstr.

19, 1191 (1958-1959).

(24)

76 Ε. Α. ZELLER 44. J. R. Fouts, On the Specificity and Mechanism of the Diamine-Diamine

Oxidase Reaction, Doctoral Thesis, Northwestern University, Evanston, Illinois, 1954; Dissertation Abstr. 14, 2194 (1954).

45. M. Silverman and E. A. Evans, Jr., J. Biol. Chem. 154, 521 (1944).

45a. H. Blaschko, Advan. Comp. Physiol. Biochem. 1, 68 (1962).

46. E . A, Zeller, in "Handbuch der allgemeinen Pathologie" ( F . Buchner, E . Letterer, and F . Roulet, eds.), Vol. II, Part 1, p. 279. Springer, Berlin, 1955.

47. H. R. Mahler, B. D . Mehrotra, and C. Sharp, Biochem. Biophys. Research Communs. 4, 79 (1961).

48. H. Tabor, Biochemistry 1, 496 (1962).

49. D. L. Keister, Some Aspects of the Metabolic Function of the Naturally Occurring Amines, Doctoral Thesis, University of Maryland Medical School, Baltimore, Maryland, 1958; Dissertation Abstr. 20, 478 (1959).

50. K. Miyaki, I. Mochida, T. Wada, and T. Kudo, Chem. Pharm. Bull.

(Tokyo) 7, 123 (1959).

51. C. C. Wu, Tai-wan i-hsiieh-hui tsa-chih 59, 862 (1960).

52. I-C. Tung, Tai-wan i-hsiieh-hui tsa-chih 59, 899 (1960).

53. H. Schmitt, P. Gonnard, and G. Glikman, Bull. soc. chim. biol. 37, 147 (1955).

54. R. W. Schayer, Proc. Soc. Exptl. BM. Med. 84, 60 (1953).

55. R. W. Schayer, Κ. Υ. T. Wu, R. L. Smiley, and Y. Kobayashi, J. Biol.

Chem. 210,259 (1954).

56. A. Pletscher and K. F . Gey, Helv. Physiol, et Pharmacol. Acta 17, C35 (1959).

57. Y. Kobayashi and R. W. Schayer, Arch. Biochim. Biophys. 58, 181 (1955).

58. Μ. E . Greig, P. H. Seay, and W. A . Freyburger, J. Neuropsychiat. 2, Suppl. 1,131 (1961).

59. Η. E . Himwich, J. Neuropsychiat. 2, Suppl. 1,136 (1961).

60. A. Burger and W. L. Yost, J. Am. Chem. Soc. 70, 2198 (1948).

61. R. E . Tedeschi, D. H. Tedeschi, L. Cook, P. A. Mattis, and E. J. Fellows, Federation Proc. 18, 451 (1959).

62. E . A. Zeller, Trans. 5th Conf. on Neuropharmacol., Princeton, New Jer­

sey, May 1959, p. 89.

63. A. R. Maas and M. J. Nimmo, Nature 184, 547 (1959).

63a. E. A. Zeller and J. R. Fouts, Ann. Rev. Pharmacol. 3, 9 (1963).

64. H. Redetzki and F . O'Bourke, Arch, intern, pharmacodyamie 130, 299 (1961).

65. H. Green and R. W. Erickson, J. Pharmacol. Exptl. Therap. 129, 237 (1960).

66. H. Green and J. L. Sawyer, J. Pharmacol. Exptl. Therap. 129, 243 (1960).

67. H. Green, S. M. Greenberg, R. W. Erickson, J. L. Sawyer, and T. Ellison, J. Pharmacol. Exptl. Therap. 136,174 (1962).

68. G. R. Pscheidt, Federation Proc. 21, 417 (1962).

69. A. Sjoerdsma, J. A. Oates, and L. Gillespie, Jr., Proc. Soc. Exptl. Biol.

Med. 103,485 (1960).

70. Μ. H. Wiseman and T. L. Sourkes, Biochem. J. 78,123 (1961).

71. H. Green and J. L. Sawyer, Proc. Soc. Exptl. Biol. Med. 104, 153 (1960).

(25)

72. N . D. Goldberg and F. E . Shideman, J. Pharmacol. Exptl. Therap. 136, 142 (1962).

73. E. A. Zeller, J. Barsky, J. R. Fouts, W. F. Kirehheimer, and L. S. Van Orden, Experientia 8, 349 (1952).

74. E. A. Zeller and J. Barsky, Proc. Soc. Exptl. Biol. Med. 81, 459 (1952).

75. A. Pletscher, K. F. Gey, and P. Zeller, Progr. in Drug Research 2, 417 (1960).

76. B. Levenberg, J. Am. Chem. Soc. 83, 503 (1961).

77. D. Bovet, Schweiz. med. Wochschr. 73, 127, 153 (1943).

78. W. Schuler and R. Meier, Helv. Physiol, et Pharmacol. Acta 13, 106 (1955).

79. J. R. Fouts, J. Barsky, and E. A. Zeller, Abstr. 126th Meeting, Am. Chem.

Soc. New York, 195U p. 45C.

80. E. A. Zeller, J. Barsky, J. R. Fouts, and J. C. Lazanas, Biochem. «7. 60, ν (1955).

81. J. Barsky, W. L. Pacha, S. Sarkar, and E. A. Zeller, J. Biol. Chem. 234, 389 (1959).

82. E . A. Zeller, J. Barsky, and E . R. Berman, J. Biol. Chem. 214, 267 (1955).

83. E . A. Zeller, L. A. Blanksma, W. P. Burkard, W. L. Pacha, and J. C.

Lazanas, Ann. Ν. Y. Acad. Sci. 80, 583 (1959).

84. S. Sarkar and E . A. Zeller, Federation Proc. 20, 238 (1961).

85. L. S. Seiden and J. Westley, Federation Proc. 21, 416 (1962).

86. M. A. Schwartz, J. Pharmacol. Exptl. Therap. 130, 157 ( 1 9 6 0 ) ; 135, 1 (1962).

87. A. N . Davison, Biochem. J. 67, 316 (1957).

88. E. A. Zeller, in "Symposium on Biochemistry and Nutrition," p. 25. Cor­

nell Univ., Ithaca, Ν . Y., 1956.

89. J. Szmuszkovicz and Μ. E. Greig, J. Med. Pharm. Chem. 4, 259 (1961).

90. F. E . Anderson, D. Kaminsky, B. Dubnick, S. R. Klutchko, W. E . Cetenko, J. Gylys, and J. A. Hart, J. Med. Pharm. Chem. 5, 221 (1962).

91. A. L. Green, Biochem. J. 84, 217 (1962).

92. Τ. E . Smith, H. Weissbach, and S. Udenfriend, Biochemistry 1, 137 (1962).

93. S. Hess, H. Weissbach, B. G. Redfield, and S. Udenfriend, J. Pharmacol.

Exptl. Therap. 124,189 (1958).

94. W. R. McGrath and A. Horita, Toxicol. Appl. Pharmacol. 4, 178 (1962).

95. K. Juva, T. Tuominen, L. Mikkonen, and E . Kulonen, Acta Pathol. Micro- biol. Scand. 51, 250 (1961).

96. D. N . Roy, S. H. Lipton, F . M. Strong, and H. R. Bird, Proc. Soc. Exptl.

Biol. Med. 102, 767 (1959).

97. E . A. Zeller, in "Histamine," Ciba Foundation Symposium, p. 339. Little, Brown, Boston, Massachusetts, 1956.

98. E . A. Zeller, J. Barsky, E . R. Berman, and J. R. Fouts, J. Pharmacol.

Exptl. Therap. 106, 427 (1952).

99. H. A. S. van den Brenk and K. Elliott, Nature 182,1506 (1958).

100. C. M. Williams, A. A. Babuscio, and R. Watson, Am. J. Physiol 199, 722 (1960).

101. M. Goldstein, A. J. Friedhoff, and C. Simmons, Biochim. et Biophys. Acta 33,572 (1959).

Hivatkozások

KAPCSOLÓDÓ DOKUMENTUMOK

An extensive literature review, based on numerous in vivo and in vitro experimental data about diterpene alkaloids in general, lead to the understanding of the structure-activ-

Ref: Katona I, Rancz EA, Acsady L, Ledent C, Mackie K, Hajos N, Freund TF. Distribution of CB1 cannabinoid receptors in the amygdala and their role in the control of

That pathogens do interfere with photosynthesis is obvious from the chlorosis they cause on many infected plants, from the necrotic lesions or large necrotic areas they produce

Előbbiek mellett az elmúlt években azonban számos olyan egyéb tényező szerepe is leírásra, jellemzésre került, amelyek elősegíthetik adott típusú, adott

48 Finally, erythrocyte Figure 1 Ischaemia and reperfusion causes the activation of platelets, which subsequently release a multitude of factors with divergent effects on infarct

All nine types of tablets were measured with the texture analyzer using the obtained, optimized settings and the in vitro DT values were compared to their in vivo

FIGURE 5 | In vivo release of serotonin (A), glutamate (B), and GABA (C) in the prefrontal cortex of mice stimulated optogenetically in their raphes (median raphe region, MRR;

The fundamental problem was that K ER and B AUGH based their assumptions solely on French ex- amples, therefore, they were writing of the French romances and not romances as