European Journal of Neurodegenerative Diseases 2022; 11(S1) July-December: S10-S14


FUNCTION AND MECHANISM OF ACTION OF ALPHA-METHYLDOPA: AN UPDATE

R.G. Bellomo *

Faculty of Physical Education Sciences, University “Carlo Bo”, Urbino, Italy.

*Correspondence to:
Rosa G. Bellomo,
Faculty of Physical Education Sciences,
University “Carlo Bo”,
61029 Urbino, Italy.
e-mail: rosa.bellomo@uniurb.it

Received: 14 November, 2022
Accepted: 30 November, 2022adobe-pdf-download-icon
2279-5855 (2022)
Copyright © by BIOLIFE
This publication and/or article is for individual use only and may not be further reproduced without written permission from the copyright holder. Unauthorized reproduction may result in financial and other penalties. Disclosure: all authors report no conflicts of interest relevant to this article.

ABSTRACT

Methyldopa is a central α-2-adrenergic receptor agonist capable of reducing blood pressure by relaxing and dilating the vessels. α-methyldopa allows blood to flow faster and for its sufficent supply to the tissues. At the level of the sympathetic nervous system, the metabolization of α-methyldopa by neurons leads to the formation of methylated norepinephrine, which is less active than norepinephrine on nerve impulses. The nerve impulses allow the release of the “false neurotransmitter”, α-methyl norepinephrine, which exerts a weaker response than the peripheral effector, an effect which is also exerted on the blood vessels. Direct stimulation with α-methyldopa produces an active, or in some cases, slightly reduced response of the post-ganglionic sympathetic system, and so methyldopa does not block the response to direct stimulation of the postganglionic sympathetic system in the vessels. In experimental animals, methyldopa has been seen to suppress vascular resistance and reduce blood pressure without affecting post-ganglionic sympathetic activity. In some neurodegenerative diseases, such as Parkinson’s disease (PD), there is a reduction in dopamine which leads to reduced mitochondrial respiratory activity and lower alpha-ketoglutarate dehydrogenase activity.

 

KEYWORDS: alpha-methyldopa, neurotransmitter, L-Dopa, CNS, Parkinson’s disease, neurodegeneration

 

INTRODUCTION

 

α-methyldopa, a methylated analogue of L-Dopa, is part of the biosynthetic pathway that leads to the synthesis of norepinephrine under physiological conditions (1). Part of the α-methyldopa is converted into α-methyl dopamine by the action of the Dopa-decarboxylase enzyme, which in turn is transformed into α-methyl norepinephrine by the β-hydroxylating enzyme (2). The antihypertensive effect of α-methyldopa is characterized by a marked decrease in peripheral vascular resistance without significant changes in cardiac output (3). This effect is accompanied by a selective reduction in renal circulation resistance with a consequent increase in glomerular flow (4). The mechanism through which α-methyldopa determines these hemodynamic effects is not yet clear and is still the subject of numerous studies and experimental investigations. From a biochemical point of view, the pharmacological action of methyldopa is different than other antihypertensive drugs and simple to understand (5). In fact, the exact correlation between the observed pharmacological effects and the biochemical events induced by this simple analogue in metabolism leading to the biosynthesis of the adrenergic neurotransmitter is not yet completely established (6).

 

Fig. 1. Chemical structure of alpha-methyldopa, a methylated analogue of L-Dopa.

 

DISCUSSION

 

In the sympathetic neuron, the metabolic use of α-methyldopa, a natural precursor L-dopa, leads to the formation of methylated norepinephrine which proves to be substantially less active than norepinephrine regarding the transmission of nerve impulses at the sympathetic level (7). For this reason, α-methyl norepinephrine is referred to as a “false neurotransmitter” (8). One of the main mechanisms of action of methyldopa is the formation of α-methyl norepinephrine, which replaces the real neurotransmitter in the deposits located at the sympathetic nerve endings (9). Upon the arrival of nerve impulses, the release of the “false neurotransmitter” results in a much milder response of the peripheral effector, i.e., with a reduction in vascular tone and the vasoconstrictor response (10). In fact, the administration of α-methyldopa in experimental animals results in a depletion of tissue deposits of norepinephrine and the appearance of methyl norepinephrine. However, the formation of the “false transmitter” and the consequent reduction of peripheral sympathetic activity does not yet fully account for the pharmacological effects induced by α-methyldopa (11).

In animals treated with α-methyldopa, the response to direct stimulation of the post-ganglia sympathetic system does not appear blocked, but only partially reduced in some cases, and even normal in others (12). Thus, methyldopa, even in large doses, does not block the response to direct stimulation of the postganglionic sympathetic system in the heart and in peripheral vascular resistance (13). For example, in dogs, it has been demonstrated that methyldopa decreases the vascular resistance of the innervated limbs without causing an appreciable decrease in post-ganglionic sympathetic activity (1).

Numerous clinical studies have also confirmed that methyldopa reduces blood pressure in humans, both in an upright or supine position (3). In these cases, the simple inhibition of the peripheral sympathetic nervous system does not seem to be sufficient to explain the mechanism of action of methyldopa. It therefore seems logical to think that other factors intervene in determining the pharmacological response to the substance. Worthy of particular interest is the hypothesis that reduced adrenergic activity in the central nervous system (CNS) is also responsible, at least in part, for the effects of methyldopa (14).

Today, the formation of methyl norepinephrine from methyldopa in the adrenergic centers of the CNS is known. It is hypothesized that the hypotensive action of methyldopa is partly due to the formation of methyl norepinephrine in the brain, which replaces the much more potent natural neurotransmitter (15). Consequently, there is a notable decrease in sympathetic activity of central origin, which in turn is responsible for the fall in peripheral vascular resistance (16). The administration of a potent peripheral inhibitor of the Dopa-decarboxylase enzyme does not prevent the hypotensive action of methyldopa. In other words, when the formation of methyl norepinephrine (“false neurotransmitter”) is blocked in the peripheral sympathetic, but not in the central adrenergic pathways, a drop in peripheral vascular resistance still occurs (17). A plausible explanation for this fact is that the formation of the “false neurotransmitter” in the CNS results in a substantial reduction in the activity of the adrenergic centers, which in turn are responsible for peripheral vascular responses (18).

This is supported by various clinical studies and experimental research. For example, in experimental animals, the infusion of methyldopa into the vertebral artery reduces systemic arterial pressure (19). The same dose injected intravenously into the peripheral circulation does not significantly change blood pressure values (20). The hypotensive effect of small doses of methyldopa injected into the vertebral artery is abolished by pre-treatment with an inhibitor of the Dopa-decarboxylase enzyme that can cross the blood-brain barrier (BBB) (19). However, if an inhibitor that does not cross the BBB is used, the action of methyldopa is not blocked. From these data it seems suggestive to conclude that an important part of the mechanism of action of the drug is carried out through a decrease in central adrenergic activity and that this decrease occurs mostly due to the increase of methyldopa in the biosynthetic chain of norepinephrine, leading to the formation of a “false neurotransmitter” such as α-methyl norepinephrine (21).

Like other antidepressant drugs, α-methyldopa has been the subject of numerous studies from which an important observation emerged, which in a certain sense, characterizes and distinguishes this drug from the others. While it has been observed that treatment with reserpine, chlorothiazide, and hydralazine generally induces an increase in renin activity in the plasma, α-methyldopa substantially reduces renin levels both in laboratory animals and in hypertensive subjects (22). In dogs, intravenous administration of α-methyldopa for 7 and 10 days causes a decrease in plasma renin activity (from 13 mg/ml to 7 mg/ml) of the produced angiotensin II (23). Again, in dogs, α-methyldopa blocks the increase in renin produced by sympathetic stimulation of the kidney without preventing the vasoconstrictor effect induced by the stimulation itself (24). Also in this case, the effect of α-methyldopa does not seem to be attributable to a generic depression or blockade of adrenergic activity, but rather either to a specific effect of the drug at the sympathetic neurological level or to an effect on the CNS (25).

In this regard, it is interesting to observe that the “false neurotransmitter” α-methyl norepinephrine synthesized by the neuron from α-methyldopa is much less active than norepinephrine in stimulating renin secretion (26). In some experiments, it has been seen that the infusion of norepinephrine in dogs induces a clear increase in plasma renin, while the infusion of equivalent doses of α-methyl epinephrine is followed by a slighter increase in renin (27). The depressant action of α-methyldopa on the renin system does not seem to be mediated by modification of the electrolyte balance or by the secretion of aldosterone (28). These parameters are not modified in hypertensive subjects after administration of the drug. Even in the clinic, it has been confirmed that α-methyldopa substantially decreases renin activity both in normotensive subjects and in non-severely hypertensive patients with renal failure (29).

From all these observations, another important characteristic of α-methyldopa emerges which certainly deserves serious consideration in the analysis of the drug’s mechanism of action. It is still unclear whether the suppression of renin activity is mediated exclusively by the action of α-methyldopa on the post-ganglionic sympathetic system, or whether it is also mediated by the action of the drug on the central adrenergic pathways (30).

 

alpha-Methyldopa and Parkinson’s Disease

In 1817, James Parkinson first described Parkinson’s disease (PD) as a disease characterized by a drastic decrease in dopamine. This neurodegenerative disease is the most widespread after Alzheimer’s disease. PD is an extrapyramidal syndrome characterized by muscle rigidity that manifests with resistance to passive movements (31). Patients affected by this neurological disorder mainly present diffuse tremors and muscle hypertonicity. The tremors usually affect the muscles of the limbs and head. In the hands, the continuous movements of the thumb and forefinger resemble pilling or rolling movements. The patient has difficulty moving and tends to fixate on their facial expression, an effect called Parkinsonian mask (32). Tremor can also occur during a state of rest and can increase in cases of anxiety and bradykinesia, causing difficulty in starting and finishing movements. The course and symptoms of PD worsen over time, even though treatment with new drugs and non-pharmacological therapies have significantly improved patients’ quality of life.

At an anatomopathological level, the disease presents progressive and prolonged neuronal degeneration of the “substantia nigra pars compacta” (33). In PD patients, the quantity of dopamine is significantly reduced, as is the concentration of neuromelanin with a consequent decrease in pigmentation (34). In addition, patients also present reduced mitochondrial respiratory activity and lower alpha-ketoglutarate dehydrogenase activity (35).

 

CONCLUSIONS

 

α-methyldopa is characterized by an antihypertensive effect and a marked decrease in peripheral vascular resistance without significant changes in cardiac output. Drastic decrease of dopamine characterizes PD which is a neurological disorder which presents muscle rigidity with resistance to passive movements. The therapies available today for PD involve dopamine replacement and other pharmacological treatments and help to prolong lifespan but are not sufficient to block the development of this disorder. Because the causes of the disease are not yet known, future studies should focus on this topic to develop therapies and improve prevention (34). In addition, the future goal of research should be to distinguish growth factors that can replace degenerated gray matter.

 

Conflict of interest

The author declares that they have no conflict of interest.

 

REFERENCES

 

  1. Kochak GM, Mason WD. The pharmacokinetics of α-methyldopa in dogs. Journal of Pharmacokinetics and Biopharmaceutics. 1985;13(4):405-423. doi:https://doi.org/10.1007/bf01061477
  2. Campbell NR, Sundaram RS, Werness PG, Van Loon J, Weinshilboum RM. Sulfate and methyldopa metabolism: Metabolite patterns and platelet phenol sulfotransferase activity. Clinical pharmacology and therapeutics/Clinical pharmacology & therapeutics. 1985;37(3):308-315. doi:https://doi.org/10.1038/clpt.1985.45
  3. White WΒ, Andreoli JW, Cohn RD. Alpha-methyldopa disposition in mothers with hypertension and in their breast-fed infants. Clinical Pharmacology & Therapeutics. 1985;37(4):387-390. doi:https://doi.org/10.1038/clpt.1985.59
  4. Grabie M, Nussbaum P, Goldfarb S, Walker BR, Goldberg M, Agus ZS. Effects of methyldopa on renal hemodynamics and tubular function. Clinical Pharmacology and Therapeutics. 1980;27(4):522-527. doi:https://doi.org/10.1038/clpt.1980.73
  5. Guillén Llera F, Reuss JM, Sagués F, Tobares N. [A comparative study of nisoldipine and alpha-methyldopa in aged patients with isolated systolic hypertension]. Revista Clínica Española. 1991;189(9):412-415.
  6. Pfeifer AK, Galambos E. Action of alpha-methyldopa on the pharmacological and biochemical effect of reserpine in rats and mice. Biochemical Pharmacology. 1965;14(1):37-40. doi:https://doi.org/10.1016/0006-2952(65)90055-9
  7. Goldberg MR, Gerkens JF, Oates JA, Robertson D. α-Methylepinephrine, a methyldopa metabolite that binds to α-receptors in rat brain. European journal of pharmacology. 1981;69(1):95-99. doi:https://doi.org/10.1016/0014-2999(81)90606-3
  8. Munion GL, Seaton JF, Harrison TS. HPLC for urinary catecholamines and metanephrines with alpha-methyldopa. The Journal of surgical research. 1983;35(6):507-514. doi:https://doi.org/10.1016/0022-4804(83)90040-9
  9. Gherezghiher T, Christensen HD, Koss MC. Studies on the mechanism of methyl-Dopa-induced mydriasis in the cat. Naunyn-Schmiedeberg’s archives of pharmacology. 1982;320(1):58-62. doi:https://doi.org/10.1007/bf00499073
  10. Frohlich ED. Methyldopa. Mechanisms and treatment 25 years later. Archives of Internal Medicine. 1980;140(7):954-959. doi:https://doi.org/10.1001/archinte.140.7.954
  11. Dhasmana KM, Spilker B. On the mechanism of l-DOPA-induced postural hypotension in the cat. British Journal of Pharmacology. 1973;47(3):437-451. doi:https://doi.org/10.1111/j.1476-5381.1973.tb08175.x
  12. Ayitey-Smith E, Varma DR. Mechanism of the hypotensive action of methyldopa in normal and immunosympathectomized rats. British Journal of Pharmacology. 1970;40(2):186-193. doi:https://doi.org/10.1111/j.1476-5381.1970.tb09912.x
  13. Baluk P, Gabella G. Some parasympathetic neurons in the guinea-pig heart express aspects of the catecholaminergic phenotype in vivo. Cell and Tissue Research. 1990;261(2):275-285. doi:https://doi.org/10.1007/bf00318669
  14. van Zwieten PA. Antihypertensive Drugs Interacting with α- and β-Adrenoceptors. Drugs. 1988;35(Supplement 6):6-19. doi:https://doi.org/10.2165/00003495-198800356-00003
  15. Nakamura K, Okada T, Ishii H, Nakamura K. Differential effects of alpha-methyldopa, clonidine and hydralazine on norepinephrine and epinephrine synthesizing enzymes in the brainstem nuclei of spontaneously hypertensive rats. Japanese Journal of Pharmacology/Japanese journal of pharmacology. 1980;30(1):1-10. doi:https://doi.org/10.1254/jjp.30.1
  16. Reid JL. Alpha-adrenergic receptors and blood pressure control. The American Journal of Cardiology. 1986;57(9):E6-E12. doi:https://doi.org/10.1016/0002-9149(86)90716-2
  17. Tayarani-Binazir KA, Jackson MJ, Fisher R, Zoubiane G, Rose S, Jenner P. The timing of administration, dose dependence and efficacy of dopa decarboxylase inhibitors on the reversal of motor disability produced by L-DOPA in the MPTP-treated common marmoset. European journal of pharmacology. 2010;635(1-3):109-116. doi:https://doi.org/10.1016/j.ejphar.2010.03.006
  18. Biaggioni I. New Developments in the Management of Neurogenic Orthostatic Hypotension. Current Cardiology Reports. 2014;16(11). doi:https://doi.org/10.1007/s11886-014-0542-z
  19. Van Zwieten PA, Mathy MJ, Thoolen MJM. Deviating central hypotensive activity of Urapidil in the cat. Journal of pharmacy and pharmacology. 1985;37(11):810-811. doi:https://doi.org/10.1111/j.2042-7158.1985.tb04973.x
  20. Ding XR, Stier CT, Itskovitz HD. Serotonin and 5-hydroxytryptophan on blood pressure and renal blood flow in anesthetized rats. The American journal of the medical sciences. 1989;297(5):290-293. doi:https://doi.org/10.1097/00000441-198905000-00004
  21. Blower PR, Poyser RH, Robertson MI. Effects of α-methyldopa on blood pressure in the anaesthetized dog. Journal of Pharmacy and Pharmacology. 1976;28(5):437-440. doi:https://doi.org/10.1111/j.2042-7158.1976.tb04650.x
  22. Moyer JH, Heider C, Pevey K, Ford RV. The effect of treatment on the vascular deterioration associated with hypertension, with particular emphasis on renal function. The American Journal of Medicine. 1958;24(2):177-192. doi:https://doi.org/10.1016/0002-9343(58)90306-1
  23. Obach R, Menargues A, Vallès JM. The pharmacokinetic profile of carbidopa in dogs. Journal of pharmacy and pharmacology. 1984;36(6):415-416. doi:https://doi.org/10.1111/j.2042-7158.1984.tb04414.x
  24. Zimmermann H, Ganong WF. Pharmacological Evidence that Stimulation of Central Serotonergic Pathways Increases Renin Secretion. Neuroendocrinology. 1980;30(2):101-109. doi:https://doi.org/10.1159/000122983
  25. Harron DWG. Distinctive Features of Rilmenidine Possibly Related to Its Selectivity for Imidazoline Receptors. American Journal of Hypertension. 1992;5(4_Pt_2):91S98S. doi:https://doi.org/10.1093/ajh/5.4.91s
  26. Fuller RW, Perry KW. Effect of uptake inhibitors on the depletion of brain norepinephrine and serotonin after alpha-methyl-m-tyrosine administration to rats. Archives internationales de pharmacodynamie et de thérapie. 1978;234(2):229-235.
  27. Rudolph CD, Kaplan SL, Ganong WF. Sites at which Clonidine Acts to Affect Blood Pressure and the Secretion of Renin, Growth Hormone and ACTH. Neuroendocrinology. 1980;31(2):121-128. doi:https://doi.org/10.1159/000123062
  28. Broughton PF, Symonas EM, Lamming GD, Jadoul FA. Renin and Aldosterone Cocentrations in Pregnant Essential Hypertensives-A Prospective Study. Clinical and experimental hypertension Part B, Hypertension in pregnancy. 1983;2(2):255-269. doi:https://doi.org/10.3109/10641958309006085
  29. Tanaka T, Seki A, Fujii J. Effect of high and low sodium intake on norepinephrine turnover in the cardiovascular tissues and brain stem of the rabbit. Hypertension. 1982;4(2):294-298. doi:https://doi.org/10.1161/01.hyp.4.2.294
  30. Garriga C, Planas JM, Moretó M. Aldosterone mediates the changes in hexose transport induced by low sodium intake in chicken distal intestine. Journal of physiology. 2001;535(1):197-205. doi:https://doi.org/10.1111/j.1469-7793.2001.00197.x
  31. Xia R, Muthumani A, Mao ZH, Powell DW. Quantification of neural reflex and muscular intrinsic contributions to parkinsonian rigidity. Experimental Brain Research. 2016;234(12):3587-3595. doi:https://doi.org/10.1007/s00221-016-4755-9
  32. Iwabuchi Y, Nakahara T, Kameyama M, et al. Impact of the cerebrospinal fluid-mask algorithm on the diagnostic performance of 123I-Ioflupane SPECT: an investigation of parkinsonian syndromes. EJNMMI Research. 2019;9(1). doi:https://doi.org/10.1186/s13550-019-0558-x
  33. Fearnley JM, Lees AJ. Ageing and Parkinson’s disease: substantia nigra regional selectivity. Brain. 1991;114(5):2283-2301. doi:https://doi.org/10.1093/brain/114.5.2283
  34. Latif S, Jahangeer M, Maknoon Razia D, et al. Dopamine in Parkinson’s disease. Clinica Chimica Acta. 2021;522(0009-8981):114-126. doi:https://doi.org/10.1016/j.cca.2021.08.009
  35. Stefano GB, Ramin R, Kream RM. Dopamine Coupling to Mitochondrial Signaling: Implications for Transplantation. Annals of Transplantation. 2016;21:35-38. doi:https://doi.org/10.12659/aot.896437

You may also like...