Next Article in Journal
Molecular Characterization of a New Tetrodotoxin-Binding Protein, Peroxiredoxin-1, from Takifugu bimaculatus
Next Article in Special Issue
Structural Requirements for the Binding of a Peptide to Prohibitins on the Cell Surface of Monocytes/Macrophages
Previous Article in Journal
Dysfunctional cGMP Signaling Leads to Age-Related Retinal Vascular Alterations and Astrocyte Remodeling in Mice
Previous Article in Special Issue
Prospects and Applications of Natural Blood-Derived Products in Regenerative Medicine
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Vasopressin and Its Analogues: From Natural Hormones to Multitasking Peptides

Department of Molecular Biochemistry, Faculty of Chemistry, University of Gdańsk, Wita Stwosza 63, 80-308 Gdansk, Poland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(6), 3068; https://doi.org/10.3390/ijms23063068
Submission received: 18 January 2022 / Revised: 24 February 2022 / Accepted: 10 March 2022 / Published: 12 March 2022
(This article belongs to the Special Issue Bioactive Peptides in Human Health and Disease)

Abstract

:
Human neurohormone vasopressin (AVP) is synthesized in overlapping regions in the hypothalamus. It is mainly known for its vasoconstricting abilities, and it is responsible for the regulation of plasma osmolality by maintaining fluid homeostasis. Over years, many attempts have been made to modify this hormone and find AVP analogues with different pharmacological profiles that could overcome its limitations. Non-peptide AVP analogues with low molecular weight presented good affinity to AVP receptors. Natural peptide counterparts, found in animals, are successfully applied as therapeutics; for instance, lypressin used in treatment of diabetes insipidus. Synthetic peptide analogues compensate for the shortcomings of AVP. Desmopressin is more resistant to proteolysis and presents mainly antidiuretic effects, while terlipressin is a long-acting AVP analogue and a drug recommended in the treatment of varicose bleeding in patients with liver cirrhosis. Recently published results on diverse applications of AVP analogues in medicinal practice, including potential lypressin, terlipressin and ornipressin in the treatment of SARS-CoV-2, are discussed.

Graphical Abstract

1. Introduction

Homeostasis of living organisms requires continuous and strict regulation. The management of all body functions involves various mechanisms and components that need to be efficiently coordinated. All the distinct organs located within the body have to cooperate and respond to the changes in the environment very quickly and efficiently. For this reason, living organisms have developed highly specialized compounds keeping the body functions under tight control. Among them are peptides and proteins that act as hormones [1,2,3] and neurotransmitters [3] that take part in the defense system [4] by controlling respiration [5], metabolism, digestion, reproduction, etc.
In healthy organisms, under physiological conditions, all components of this complex network cooperate undisturbed. Pathological states, when some elements of this system fail, may result in various diseases, e.g., resistance to insulin results in Diabetes mellitus type 2 [6,7], while excess and aggregation of amyloid β peptide has a role in Alzheimer’s disease progress [8,9]. Currently, when the knowledge about the role and structure of biologically active peptides is pretty advanced, they are considered as potentially attractive therapeutics. Peptide synthesis is well developed [10,11] and peptide drug market is rapidly expanding [12,13]. Peptides occur naturally and are usually considered to be safer than synthetic drugs. They are also more selective and specific and, what is particularly important, they are degraded into nontoxic metabolites (amino acids) [14,15]. Taking it all into account, the introduction of peptides as drugs should be easy to handle.
However, despite their involvement in various bioactivities in living organisms and quite extensive knowledge regarding synthesis and structure–activity relationships, the utilization of peptides as drugs is not so straightforward [13,16]. In addition to indisputable advantages, peptides are not free from significant drawbacks [5,17]. The first major obstacle is problematic oral delivery, as most peptide drugs are characterized by low stability in the gastrointestinal tract. The short half-life of peptides is explained by the presence of peptidases in the body causing their degradation. The hydrophilic nature of peptides results in poor intestinal permeability [18,19] and, consequently, insufficient oral bioavailability. Currently, most peptide drugs are applied parenterally, which is normally considered as inconvenient. Oral application seems to be the easiest and most favorable way of taking medications [18]. Unfortunately, these are not the only problems with applying peptides in treatment. Their administration far from the target brings more biological barriers that drastically decrease peptides’ oral bioavailability.
Even though the utilization of peptides as drugs is challenging, their valuable properties make them extremely attractive candidates to become therapeutics. For this reason, modifications of peptide-based drug candidates increasing their metabolic stability and bioavailability, and retaining their functions are of high demand. Several significant achievements have been made, such as the introduction of non-proteinogenic amino acids, cyclization, modifications of termini or peptide bonds (pseudopeptides) and the design of peptide mimetics [13,15,20].
An obvious example of naturally occurring peptide with an especially valuable function in the body is vasopressin (AVP). This neurohormone is synthesized in overlapping regions in the hypothalamus in large magnocellular neurons situated in supraoptic nuclei and paraventricular nuclei [21,22]. In a healthy organism, AVP is responsible for the regulation of plasma osmolality by maintaining fluid homeostasis, e.g., it induces water reabsorption from urine and brings it back to circulation [23,24,25]. It is also a vasoconstricting agent taking part in blood pressure increase. In medical practice, it is used mostly in the treatment of vasodilatory shock; in septic shock during sepsis; and in the treatment of hypertension, edema and congestive heart failure [26]. After its discovery and subsequent chemical synthesis in 1953 [27], the next natural step was the urge to decipher its exact mechanism of action. For this reason, various AVP analogues, acting both as agonists or antagonists of its receptors, were designed and investigated. This review aims at presenting the most relevant ones. Figure 1 presents the milestones in the development of key peptide analogues of AVP, both of natural and synthetic origins.

2. Vasopressin

Vasopressin (arginine vasopressin, argipressin, AVP, Figure 2a) is a nonapeptide comprising a tripeptide tail and a cyclic structure formed by six residues with a disulfide bridge between cysteines at positions 1 and 6 [36,37,38]. It is found exclusively in mammals and was isolated and synthesized for the first time in 1953 by Vincent du Vigneaud et al., who received the Nobel Prize in 1955 [27,39]. Its chemical structure is similar to another peptide hormone, oxytocin (OXT), which is the hydrophilic, cyclic nonapeptide (Cys(&)-Tyr-Ile-Gln-Asn-Cys(&)-Pro-Leu-Gly-NH2, Figure 2b) [38,40,41] composed of a six-membered ring with a disulfide bridge and tripeptide tail. The main differences are at positions 3 (Phe (AVP) → Ile (OXT)) and 8 (Arg (AVP) → Leu (OXT)) [42]. OXT was found in mammals in 1906. It is synthesized as inactive protein precursor of OXT [41] and active OXT is released from the posterior pituitary gland. OXT is also found in multiple organs, including heart, placenta, uterus and testes [40]. Its key role is the modulation of social bonding, emotions, maternal behavior, etc. [41,43]. AVP is synthesized from a precursor protein named preprovasopressin, which consists of AVP, neurophysin II and copeptin [44]. They are packed into neurosecretory vesicles and transported to the nerve endings located in neurohypophysis [27,45].
AVP binds to three G protein-coupled receptors: V1a (Avpr1a or V1), V1b (Avpr1b or V3) and V2 (Avpr2) [46]. They all share a high primary sequence homology, but have great diversity in their functional properties (Figure 3) [47]. V1a is a vascular receptor found primarily in vascular smooth muscle cells but was also discovered in platelets, liver, blood vessels, renal mesangial cells, brain, uterus and adrenal cortex [48,49,50,51]. Activated V1a receptor controls blood pressure, glycogenolysis and contraction of mesangial cells. It also facilitates vasoconstriction through the activation of phospholipase A, C and D, which results in the mobilization of intracellular calcium ions (Ca2+) [21,49] and leads to platelet aggregation (procoagulant property) [52]. Myocardial infarction and stress increase V1a receptor expression, while the activation of V1a in the brain can exacerbate brain edema [49]. V1b receptor is not only expressed in the anterior pituitary [21,52] but also in peripheral tissues such as the pancreas, thymus, lungs, heart and kidneys. It controls the excretion of an adrenocorticotropic hormone (ACTH) [21,23,53] and regulates cardiovascular and adrenal functions by mediating catecholamine excretion. It also participates in the modulation of behavior in social interactions and in the memory process [21]. Together with V1a, it activates the phosphatidylinositol hydrolysis pathway (formation of 1,2-diacylglycerol and inositol-1,3,4-triphosphate) and mobilizes intracellular Ca2+ [49,54]. AVP receptor V2 is found in the kidneys where it is expressed on basolateral surface of renal tubular cells [45,52,55,56]. The main role of activated V2 receptor is water reabsorption from the urine, i.e., reduction in urine volume (antidiuretic effect) [21,26,49]. The mechanism of water reabsorption is based on changes in plasma osmolality [57]. This results in electrical simulation (depolarization of cells) and the production of AVP [21,27]. AVP binds to V2 receptor expressed in the basolateral side of principal cells responsible for the regulation of water permeability in collecting duct of the kidney. This activates the adenylate cyclase and the production of cyclic adenosine monophosphate (cAMP). The latter further activates protein kinase (PKA) that phosphorylates some of the C-terminal residues of aquaporin-2 (AQP-2). In effect, AQP-2 is inserted into the apical side of principal cell-forming AQP-2 water channels. Water is reabsorbed via these channels into the principal cells and exits through AQP-3 and AQP-4 water channels to the interstitium. In this manner, water is reabsorbed from urine into the blood [22,45,52,58,59,60,61]. In addition to this classic PKA/cAMP-dependent pathway, cAMP functioning might also be mediated by the exchange of protein directly activated by cAMP [62]. The translocation of AQP-2 from the basolateral to apical regions of cells is dependent on AVP-induced Ca2+ mobilization. Thus, Ca2+ mobilization is crucial for water permeability in the inner medullary collecting duct. This process may be blocked with intracellular Ca2+ chelation with 1,2-bis(2-aminophenoxy)ethane-N,N,N,N-tetraacetic acid (BAPTA) [63]. The process of plasma osmolality increase/decrease is so sensitive that even a change of 1% can alter AVP release/inhibition [57]. After the release of AVP, only 10–20% of the total amount is present in blood circulation and its plasma half-life is 6–20 min [24,52,64]. The lack of AVP results in diabetes insipidus, which means that no other hormone can replace AVP [24]. The expression of V2 was also detected in central nervous system (CNS) where its presence was associated with anxiety, development, alcohol preference and aggression [26,49].
Release of AVP can be induced under all types of stress—immunological, physical and emotional [49,65]. The concentration and secretion of AVP may be regulated by insulin [45], changes in plasma osmolality (decrease in serum osmolality reduces AVP secretion, while an increase in serum osmolality promotes the secretion of AVP) [57,66] or use of morphine, alcohol and nicotine [57]. Depending on the various external factors influencing the body and its actual internal needs, AVP is able to regulate numerous biological processes. They include the following: increase in cortisol concentration through the induction of corticotropic axis stimulation [52], insulin secretion, release of coagulant factors [45], regulation of lipid metabolism [56], stimulation of water reabsorption, regulation of glucose mechanism, platelet aggregation, modulation of neuronal functions [54], growth of surface placental leucine aminopeptidase (P-LAP; enzyme crucial to maintain pregnancy) level [67], rise of plasma osmolality, regulation of blood pressure and volume [66] and reduction in arrhythmias and tachycardia [26]. AVP has also been shown to affect modulation of fear and memory [27], maintain cardiovascular homeostasis through vascular smooth muscle cell contraction [68], regulate pancreatic islet function [46], reduce hemorrhage [69] and, together with angiotensin II, preserve fluid and electrolyte balance [70]. Its expression in CNS has an impact on a variety of brain functions such as social interaction and recognition, aggression, maternal behavior, pair-bonding behavior and depression [21]. Although its role is irreplaceable, AVP frequently induces severe side effects because of its systemic vasoconstricting effect. This, in turn, induces an increase in peripheral resistance, with a reduction in cardiac output and heart rate and ultimately of coronary blood flow. These hemodynamic effects may result in myocardial ischemia or infarction, cardiac arrhythmias, mesenteric ischemia, ischemia of the limbs and cerebrovascular accidents [71]. As far as the trade market is concerned, AVP is marketed under different names in different countries worldwide. In the Unites States of America, it is used under the international name Pitressin or Pressyn [27], while in Spain it can be found under the name Vasostrict [72]; in India as Cpressin, Cevas-40, Cpressin P, etc. [73]; in Germany as Empressin [74]; and in France under the name Reverpleg [75].

2.1. Vasopressin Analogues

Despite its significant role in the treatment of various diseases, such as diabetes insipidus [47], vasodilatory shock and hypertension [26], the use of AVP is associated with some serious limitations, including short biological half-life, lack of specificity for V1 and V2 receptor and side effects [71,76]. AVP can often induce hyponatremia, decrease cardiac output and platelet count or increase the direct and indirect bilirubin level in the blood. The efforts to solve some of these shortcomings and discover new analogues characterized by better activity and selectivity [38] constitute a very important area of research. In 1980, AVP receptors (V1a, V1b and V2) were cloned and characterized [21], which was a milestone in the development of new AVP analogues. In the last forty years, various peptide and non-peptide analogues, acting as agonists or antagonists of these receptors, were designed and synthesized [47]. Nevertheless, analogues that could be very active, potent and truly selective for AVP receptors still remain an area of great interest. Such analogues constitute a pharmacological toolbox essential to effectively study the mechanism of action of AVP and ligand–receptor interactions [77].

2.1.1. Non-Peptide Synthetic Analogues of AVP

The synthesis of small molecules mimicking larger compounds is of great interest and a great challenge in chemistry, especially when it comes to non-peptide molecules intended to imitate peptides. In the last few years, some low molecular weight AVP analogues were synthesized and presented good affinity to AVP receptors. Such non-peptide agonists and antagonists of AVP receptors are presented in Figure 4. Compound 1 had a high affinity to human V1a receptor (hV1a) and potently inhibited AVP-induced physiological responses of human coronary artery smooth muscle cells (CASMC) [68]. Compounds 2 and 3 showed high binding efficiencies to the hV1a receptor [78] and a similar chemical structure based on 4,5-diphenyl-1,2,4,-triazole moiety. Interestingly, compound 4 revealed high in vitro hV1a antagonistic affinity, while compound 5 had reduced hV1 binding affinity. This suggests that α-methyl substitution shifted the binding mode of the adjacent phenyl ring [79]. Compound 6 was a potent and selective brain-penetrant hV1a antagonist as well as 7 (belovaptan), which demonstrated an amelioration of socialization and communication [80]. Belovaptan underwent four clinical tests for use in the treatment of autism spectrum disorder, but only one was completed, while two were terminated and one was withdrawn [81]. Analogue 8 resulted in good bioavailability and extremely strong binding to the rat V1a and hV2 receptor. It had a potent diuretic effect and showed antagonism of AVP-induced hypertension [82]. Hydrochloride 9 showed high selectivity for rat V1a/V2 and hV1a/hV2 receptors and was under the clinical testing for patients with edema [83,84]. Compounds 10 and 11 were V2 receptor agonists characterized by high binding activity [50,61].

2.1.2. Natural Peptide Analogues of Vasopressin

Lypressin

Lypressin (lysine vasopressin, LVP; Figure 5a) differs from AVP by one amino acid, i.e., Lys instead of Arg at position 8 (Table 1) [64]. The disulfide stretching bands in the Raman spectra of LVP exhibited small shoulders, which indicates the presence of some conformational flexibility in the disulfide moiety. Likewise, the circular dichroism spectra reveal the presence of more than one conformation of the disulfide unit [85].
LVP is the porcine antidiuretic hormone [86], and its plasma half-life is 5–7 min [64]. Disulfide bridge may be reduced, which leads to a loss of biological activity [87]. LVP is used to treat diabetes insipidus and to improve vasomotor tone and blood pressure [86]. It is available for clinical use as a nasal spray [88]. Such an application form exerts its efficiency up to eight hours, and patients treated with it are free of allergic reactions. LVP does not cause any elevation of blood pressure and has been proved to be safe for use during pregnancy and parturition [89].
Table 1. Comparison of AVP with its natural peptide analogues. Amino acid residues that are not present in AVP are marked in bold.
Table 1. Comparison of AVP with its natural peptide analogues. Amino acid residues that are not present in AVP are marked in bold.
AnalogueSequenceSourceMain ApplicationRefs.
Arginine vasopressin, argipressin, AVPCys(&)-Tyr-Phe-Gln-Asn-Cys(&)-Pro-Arg-Gly-NH2human and other mammalsantidiuretic effect, maintenances cardiovascular homeostasis, increases blood pressure in septic shock[38,44,45,46,49,65]
Lysine vasopressin, lypressin, LVPCys(&)-Tyr-Phe-Gln-Asn-Cys(&)-Pro-Lys-Gly-NH2pigsantidiuretic agent, hemostatic, vasoconstrictor agent[64,86,87]
PhenypressinCys(&)-Phe-Phe-Gln-Asn-Cys(&)-Pro-Arg-Gly-NH2marsupials (gray and red kangaroo, tammar and quokka wallaby)increases the reabsorption of water in the kidneys and blood pressure[90,91]
Since 1960, many LVP analogues have been synthesized. The effect of modifications in the side chains of various amino acids on the biological activity has been examined. The most relevant analogues and the results of their biological investigation, including antidiuretic and vasopressor activities, are presented in Figure 6 and in Table 2.
Jarial et al. [92] compared the diagnostic accuracy of LVP and human corticotrophin releasing hormone) as stimulating agents for ACTH release during bilateral inferior petrosal sinus sampling (BIPSS) to localize and lateralize the source of ACTH in patients with Cushing’s syndrome (CS). The results showed that LVP stimulating ACTH secretion in BIPSS test for CS and 10 units LVP correctly lateralized the pituitary adenoma in three-fourths of patients. BIPSS was suggested to be used as a good method to identify ACTH-secretion sources [93].
Table 2. Peptide LVP analogues and their biological activity.
Table 2. Peptide LVP analogues and their biological activity.
PeptideAntidiuretic Activity
Units/mg
Vasopressor Activity
Units/mg
Oxytocic Activity
Units/mg
Other Activities and Comments
LVP 203 ± 7 [94]
240 ± 13 [95,96]
243 ± 3 [94]
266 ± 18 [96]
7.3 ± 0.2 [96]
4.8 ± 0.3 [94]
AVD a = 48 ± 2 units/mg [94]
dLVP (12) b301 ± 11 [96]
550 ± 1.7 [97]
126 ± 2 [96]
145 ± 7 [97]
12 ± 0.5 [96,97]-
[Dbt2]dLVP (13) cnr dnrnrAVD = pA2 = 6.97 e; M = 1.1 × 10−7; no measurable antagonistic, oxytocic and pressor activities [98]
[Tyr(OMe)2] LVP (14)1.5–3 [99]79 [99]antioxytocic and antipressor properties, antagonistic character of these analogues results from the bulky, lipophilic substituents on the aromatic ring rather than from the blocking or elimination of the phenolic group [99]
[Tyr(OEt)2] LVP (15)nr5 [99]
[Tyr(OX)2] dLVP (1619) f0.5–2.0 units/μmol [100]0.5–3.0 units/μmol [100]weak agonistic properties; in the rat, none of the analogues inhibited the antidiuretic action of LVP when the two substances were administered together in a single injection; completed inhibition was obtained when X = Et; antagonistic potency decrease with increasing size of alkyl substitution [100]
[Thi3]LVP (20)332 ± 32 [101,102]243 ± 5 [101,102]19 ± 0.5 [101,102]AVD = 87 ± 4 units/mg; steric size in position 3 plays significant role in the manifestation of vasopressin-like activities [101,102]
[Ile3]LVP (21)24 ± 3 [101,102]130 ± 13 [101,102]78 ± 10 [101,102]AVD = 210 ± 3 units/mg [101,102]
[Ser3]LVP (22)~0.08 [102]<0.01 [102]nruterotonic activity ≤ 0.01 units/mg [102]
[Tyr3]LVP (23)0.18 [102]1.6 [102]
[diHPhe3] LVP (24)125–130 [102]129–132 [102]uterotonic activity = 6 units/mg; effective agonist; position 3 is not very restrictive for vasopressor receptors and antidiuretic potency [102]
[Leu4]LVP (25)1–2 [95]1.33 [95]negligible [95]AVD = 1 unit/mg [95]
[Leu4]dLVP (28)5–6 [95]
10.5 ± 0.9 [97]
0.55 [95]
0.9 ± 0.1 [97]
0.054 ± 0.002 [97]AVD = 4.60 units/mg [95] high affinity for the rat V1b receptor, very low affinities for the rat V1a and V2 receptor, potent agonists for the rat V1b receptor, weak agonists for the rat antidiuretic activity [97], Gln is not essential for biological activity [95]
[Abu4]LVP (26)707 [95]nrnrGln is not essential for biological activity [95]
[Abu4]dLVP (29)729 [95]3.5 [95]
[Ala4]LVP (27)707 ± 107 [96]10.2 ± 0.6 [96]1.54 ± 0.1 [96]-
[Ala4]dLVP (30)729 ± 26 [96]3.5 ± 0.2 [96]1.51 ± 0.05 [96]
[Cha4]dLVP (31)0.82 ± 0.01 [97]0.043 ± 0.008 [97]pA2 = 6.48 ± 0.03
M = 3.41 × 10−7 ± 0.2 [97]
high affinity for the rat V1b receptor, very low affinities for the rat V1a and V2 receptor, potent agonists for the rat V1b receptor, weak agonists for the rat antidiuretic activity [97]
[Orn4]dLVP (32)7.8 ± 0.4 [97]0.23 ± 0.02 [97]3.1 ± 0.1 [97]-
[Arg4]dLVP (33)784 ± 54 [97]83 ± 4 [97]0.15 ± 0.02 [97]
[diMeGln4] LVP (34)1.88 ± 0.04 [94]1.27 ± 0.03 [94]<0.05 [94]AVD ≤ 0.1 units/mg [94]
[Ala5]LVP (35)~0.2 [96]0.15 ± 0.01 [96]<0.001 [96]carboxamide group is essential for activity [96]
[Ala5]dLVP (36)~0.05 [96]~0.015 [96]<0.002 [96]
[diMeAsn5] LVP (37)5.5 ± 0.3 [103]2.55 ± 0.05 [103]<0.05 [103]AVD = 0.39 ± 0.03 units/mg; hydrogen atoms of carboxamide group are not essential for antidiuretic activity [103]
[Lys(N-Gly)38]dLVP (38)nrnrnrmore powerful and prolonged analgesia compared to LVP [104]
[Eda9]LVP (39)<0.05;
0.002 [105]
-
[Eda9]dLVP (40)126 [105]
a AVD = avian vasodepressor; b d = deamino cysteine (Cys1); c Dbt = 3,5-dibromo-l-tyrosine; d nr = not reported; e pA2 values in vitro represent the negative logarithm to the base 10 of the average molar concentration (M) of the antagonist that reduces the response to 2x units of agonist to equal the response seen with x units of agonist administered in the absence of the antagonist; f X = ethyl- (Et), propyl- (Pr), butyl- (Bu), hexyl- (Hex).

Phenypressin

Phenypressin ([Phe2]AVP, Figure 5b) is another neuropeptide belonging to the vertebrae vasopressin family [90]. It has been found mostly in some species of the Macropodidae family, particularly in eastern gray kangaroo, red kangaroos, tammar wallaby and the quokka wallaby [31,34,91,106]. Phenypressin, similarly to AVP, is synthesized in the hypothalamus and travels to the posterior pituitary and is then released into the vesicles. Phenypressin was found to be less abundant than vasopressin-like peptides in marsupials [106]. The name phenypressin was suggested by Chauvet et al. [34] because Tyr2 in AVP is replaced by Phe (Table 1) and was synthesized for the first time in 1962 by Huguenin and Boissonnas [31]. The comparison of the pharmacological properties of phenypressin with those of AVP revealed a reduction in rat antidiuretic activity from 435 to 375 units/µmol and rat blood pressure activity from 435 to 130 units/µmol [107].

2.1.3. Synthetic Peptide Analogues of Vasopressin

Numerous peptide analogues of AVP have already been synthesized. They are based mostly on the changing of one or more amino acids in its sequence or side chain modification such as N-methylation, α-alkylation and local or global cyclization. Extensive research on AVP analogues allowed the determination of the role of specific residues in AVP biological activity. Table 3 presents selected analogues of this hormone described in this section together with their vasopressor, antidiuretic and oxytocic potencies.
Tyr at position 2 is considered to contribute to the initiation of the pressor response to AVP [51,108]. Indeed, modification in Tyr2 of AVP with methyl group (Me) ([Tyr(OMe)2]AVP], 41, Figure 7) resulted in an analogue with decreased vasopressor potency and higher antidiuretic activity [109]. The acetylation of its free amine group of Cys ([Cys1(N-Ac),Tyr(OMe)2]AVP, 42, Figure 7) reduced antidiuretic activity 15,000 times [110]. Tyr 2 replacement with 3,3-diphenyl-l-alanine ([diPhe2]AVP, 43, Figure 7) or its d-enantiomer (44, Figure 7) increases antidiuretic activity, which is even more remarkable in the case of [d-diPhe2]AVP. Vasopressor potency of both analogues was completely abolished [111,112]. Modification at position 2 with 2-aminoindane-2-carboxylic acid (Aic, [Aic2]AVP, 45, Figure 7) did not alter the interaction with the receptor and preserved antidiuretic activity similar to AVP [77], while substitution of position 2 with 1-aminocyclopentane-1-carboxylic acid ([Apc2]AVP, 46, Figure 7) resulted in higher antidiuretic activity [56].
Phe at position 3 is important for binding to the receptor and determines hormone antidiuretic and uterotonic activity [51]. Its replacement with l-1-naphthylalanine (l-1-Nal), inversion of configuration at position 8 and deamination of the N-terminus yield the [l-1-Nal3, d-Arg8]dAVP, 47, Figure 7) potent agonist of V2 receptor. Interestingly, when l-1-Nal in position 3 is replaced by l-2-Nal (48, Figure 7), the selective V1 receptor antagonist is obtained [113].
Pro7 has a key role in the orientation of tripeptide tails and in the conformational constraints of peptide and binding to V2 receptor [114,115]. Analogues with Pro7 replaced by l-sarcosine ([Sar7]AVP, 49, Figure 7) or N-methyl-l-alanine ([NMeAla7]AVP, 50, Figure 7) turned out to be specific antidiuretic agents, while their vasopressor activity was abolished. They also retained high binding affinities relative to renal vasopressin receptors V2, but the binding affinity to V1 receptor was lost [114]. Deletion of Pro retained antidiuretic antagonist activity. Pro7 was reported to be important for binding and activation of the V2 receptor, whereas it was not necessary for V2 receptor antagonistic activity [115].
To examine the importance of the basic amino acid at position 8 relative to AVP and LVP biological activity, the analogue with l-homonorleucine (HNle) was synthesized [116]. The results showed that [HNle8]AVP (51, Figure 7) has 8% of LVP activity. Its antidiuretic activity in ethanol-anesthetized rats equals about 10 units/mg, as compared to about 250 units/mg for LVP. These results indicate that amino group in position 8 is not indispensable in terms of showing vasopressin-like activities, although a cation-forming group is needed for potencies comparable to those of the natural hormone. Interestingly, any basic residue in this position is not sufficient for high potency. This was confirmed by the example of [His8]AVP (52, Figure 7), which showed about 270-fold and 130-fold decrease in pressor activities when compared to AVP and LVP, respectively [117]. The relative basicity of amino acid residue has a considerable influence on the degree of pressor activity. To examine if the hormone is bound to the receptor and not through an ionic bond but by a hydrogen bond, AVP, an analogue with l-hydroxynorleucine [HyLeu8]AVP, 53, Figure 7) in position 8 was synthesized. Its hormonal activities are moderate but were significantly higher than those of [HNle8]AVP. This suggests that the hydrogen bond was able to substitute, to some extent, the ionic bond between hormone and receptor. Alternative explanations, such as polar interactions or hydrophilic effect, cannot be ignored [116].
Table 3. Synthetic peptide AVP analogues and their biological activity.
Table 3. Synthetic peptide AVP analogues and their biological activity.
PeptideAntidiuretic Activity
Units/mg
Vasopressor Activity
Units/mg (for Agonist)
pA2 (for Antagonist)
Oxytocic Activity
Units/mg (for Agonist)
pA2 (for Antagonist)
AVP
(Figure 2a)
323 [118]
450 (t1/2 = 60)
450 (t1/2 = 200) [119]
369 [118]
412 [119]
14 [118]
[Tyr(OMe)2]AVP
(41)
386 ± 36 [109]pA2= 9.7 ± 0.5 a [109]pA2 = 7.44 ± 0.12 (no Mg2+) and 6.34 ± 0.19 (in 0.5 mM Mg2+) [109]
[Cys1(N-Ac), Tyr(OMe)2]AVP (42)0.026 ± 0.002 [110]pA2= 7.18 ± 0.08 [110]pA2 = 7.29 ± 0.08 (no Mg2+) and 6.73 ± 0.14 (in 0.5 mM Mg2+) [110]
[diPhe2]AVP (43)450 (t1/2 = 60)
9000 (t1/2 = 200) [112]
0 [112]pA2 = 7.00 ± 0.20 [112]
[d-diPhe2]AVP (44)1000 (t1/2 = 60)
45,000 (t1/2 = 200) [111,112]
0 [111]pA2 = 7.82 ± 0.39 [111]
[Aic2]AVP (45)450 (t1/2 = 60)
45,000 (t1/2 = 200) [77]
9.4 ± 2.8 [77]pA2 = 7.27 ± 0.22 (no Mg2+) [77]
[Apc2]AVP (46)1800 (t1/2 = 60)
1800 (t1/2 = 200) [56]
13.4 ± 3.8 [56]0.2 units/mg
pA2 = 6.0 (no Mg2+) [56]
[l-1-Nal3, d-Arg8]dAVP (47)2.2 ± 0.83 [113]nr bnr
[l-2-Nal3, d-Arg8]dAVP (48)3.79 ± 1.31 [113]nrweak [113]
[Sar7]AVP (49)188 ± 19 [114]3.6 ± 0.2 [114]nr
[NMeAla7]AVP (50)343 ± 54 [114]10.6 ± 0.4 [114]nr
[HNle8]AVP (51)10 [116]21.4 ± 1.0 [116]nr
[His8]AVP (52)nr1.5 [117]nr
[HyLeu8]AVP (53)70 [116]30 [116]nr
a pA2 values in vitro represent the negative logarithm to the base 10 of the average molar concentration (M) of the antagonist that reduces the response to 2x units of agonist to equal the response seen with x units of agonist administered in the absence of the antagonist; b nr = not reported; pA2 values represent the negative logarithm to the base 10 of the average molar concentration of an antagonist that reduces the response to 2x units of the agonist to x units of the agonist.

Desmopressin

Desmopressin (1-desamino-8-d-Arginine vasopressin, 1-deamino-8-d-arginine vasopressin, dDAVP; 54, Figure 8) is a synthetic, deaminated analogue of AVP, bearing the substitution in position 8, where d-Arg replaces l-Arg (Table 4) [120,121]. As compared to AVP, dDAVP has a longer plasma half-life of 90–190 min [64] and is more resistant to degradation by pancreatic proteases (mostly by trypsin) and has superior affinity to V2. dDAVP is referred to as the first generation analogue of vasopressin. The first chemical synthesis of dDAVP was conducted in a solution in 1967 [122,123]. The introduction of solid phase peptide synthesis (SPPS) strategy enabled its production on a bigger, commercial scale [124].
dDAVP was subjected to several experiments assessing structure–activity relationships (SARs). Nearly all amino acid residues were successively substituted with Ala, except both Cys residues located at positions 1 and 6, which are responsible for cyclization. It was demonstrated that N-terminal amino acids located at the peptide’s loop play a crucial role in antiproliferative activity against the aggressive MDAMB231 human breast cancer cell line [139]. Substitutions in the positions 2, 3 and 5 caused the highest reduction (50–60%) in cytostatic effects. Such an observation is in accordance with a molecular modelling study [140] and other previous research demonstrating the pivotal role of vasopressin’s N-terminal end in binding with V2 receptor [141,142]. The substitution of Gln4 and Gly9 was more tolerable and caused a 30% reduction in activity. On the other hand, the replacement at C-terminal position 7 (Pro replaced either with Ala or (2S,4R)-p-hydroxyproline (Hyp)) and 8 (d-Arg) did not affect the antitumor effect.
As shown by Sawyer et al. [143], a single substitution at position 4 (replacement of Gln with Val, 59, Figure 9) resulted in dDAVP analogue with undetectable vasopressor action and antidiuretic activity about four times that of AVP. As compared to dDAVP, [Val4]dDAVP had about 10-fold higher affinity, expressed as the Ki value, to human V2 receptor. However, similarly to dDAVP, it was also a potent agonist of the human V1b receptor [141]. It was demonstrated recently that [Val4]dDAVP could rescue the function of the N321K-mutated V2 receptor without significant side effects on the V1-initiated vasoconstriction [144]. [Val4]dDAVP had a similar potency as AVP to stimulate the production of cAMP by this mutant receptor but, in contrast to the natural hormone, it did not promote vasoconstriction of peripheral mouse arterioles. Thus, the administration of [Val4]dDAVP may be considered as beneficial for patients with nephrogenic diabetes insipidus. The substitution of Tyr2 with bulky, sterically restricted Aic ([Aic2,Val4]dDAVP, 60, Figure 9) resulted in peptides with high antidiuretic activity comparable to that of dDAVP [66]. Similarly, the introduction of bulky diPhe ([diPhe2,Val4]dDAVP], 61, Figure 9) or d-diPhe in position 2 of [Val4]dDAVP generated exceptionally potent antidiuretic agents in rats with significantly prolonged activities [112].
Another [Val4]dDAVP analogue, in which Asn5 was replaced with Gln (62, Figure 9), has been studied extensively as a potential anticancer agent [145]. It was demonstrated that [Val4,Gln5]dDAVP had significant antiproliferative activity against V2-receptor expressing MCF-7 human breast carcinoma cell line [145,146]. As compared to dDAVP, [Val4,Gln5]dDAVP exerted superior inhibitory effect on breast cancer cell proliferation and colony formation, reduced tumor growth and angiogenesis and improved the survival rate in the triple-negative MDA-MB-231 xenograft in nude mice [147]. Treatment with [Val4,Gln5]dDAVP caused a complete inhibition of metastatic progression in hormone independent and metastatic F3II breast cancer mouse model [148]. The combination of [Val4,Gln5]dDAVP either with chemotherapeutic agent paclitaxel in treatment of MDA-MB-231 tumor bearing nude mice or with carmustine in metastatic F3II breast cancer model resulted in greater tumor growth inhibition as compared to single-drug regimen [147]. Moreover, the beneficial effect involved a reduction in local aggressiveness and impairment of both tumor invasion and infiltration of the skin. [Val4,Gln5]dDAVP reduced in vitro proliferation and migration in aggressive V2-receptor expressing human lung (NCI-H82) and prostate (PC-3) cancer cell lines with neuroendocrine characteristics [149]. In contrast to dDAVP, treatment with [Val4,Gln5]dDAVP abolished the formation of experimental metastases in the lungs of mice [148]. [Val4,Gln5]dDAVP impaired the spread and growth of colorectal cancer cells in the liver and reduced experimental lung colonization [150]. Importantly, studies employing animal models revealed that [Val4,Gln5]dDAVP is safe and, similarly to dDAVP, is not toxic [148]. The presence of Val4 in place of Gln4 offers higher hydrophobicity, whereas Gln introduced in position 5 is considered as less susceptible to deamidation than Asn. The relevance of chirality at position 4 was confirmed. Moreover, it was shown that tetrapeptides corresponding to the conformational loop of dDAVP (fragment 2–5 flanked with Cys residues) display similar antiproliferative effects on MCF-7 cell cultures as longer, parental peptides [145].
Wiśniewski et al. [151] described the quest for novel, potent and short-acting V2 receptor agonists. Interestingly, the reduction in antidiuretic duration may be of particular interest in older people with impaired function of kidneys, which are responsible for the elimination of dDAVP via passive glomerular filtration into urine. The prolonged antidiuresis may result in adverse effect known as hyponatremia (serum sodium concentration lower than 135 mmol/L). A series of C-terminally truncated analogues of dDAVP deprived of Gly9 amide and modified at positions 2, 3, 4 or 7 was analyzed. In some of them, the disulfide bridge was replaced with carba-thioether linkage. In order to improve systemic clearance and increase receptor selectivity, more lyophilic peptides were designed. Some novel agonists retained potent in vitro human V2 receptor activity and presented substantially improved selectivity versus the related human V1a, V1b and oxytocin receptors. The most promising results were obtained for two peptides having (β-(4-chlorophenyl) alanine ([Tyr(4-Cl)2]dDAVP, 63, Figure 9) instead of Tyr2, β-(2-thienyl) alanine ([Thi3,Val4]dDAVP, 64, Figure 9) in place of Phe3 and Val in position 4. Both have carba-thioether modification of the disulfide bridge. Agmatine was connected to one of the peptides’ C-terminal end. Both novel analogues produced dose-related decreases in urine output comparable to that of dDAVP. They also presented shorter duration of antidiuretic action due to their higher systemic clearance and shorter half-lives. It was concluded that a higher number of carbon atoms in the structures (thus higher lipophilicity) resulted in increased systemic clearance. It is worth noting that some compounds were cleared exclusively via non-renal mechanisms, presumably proteolysis and transport into organs [151].
dDAVP may be administered at various doses by different routes, including intravenous and subcutaneous injections (the most common single dose for a human is 2 μg), an intranasal spray (10–20 μg) and drop, orally available solid tablets (mostly between 200 and 400 μg) or sublingual melt formulation (60–240 μg, oral lyophilizate/orally disintegrating tablet) [152]. The last route is becoming more popular because orally administered lyophilizate melted under the tongue is fast dissolving, has a higher rate of bioavailability than a tablet and needs shorter time periods to reach maximal biological effect [153,154,155,156]. Lottmann et al. showed that treatment with sublingual dDAVP, which excludes the need to swallow tablets, is related to higher compliance with children aged 5–11 years [157]. Increased patient conformity (aged 5–15 years) was also documented by Juul et al. [155]. The bioavailability of dDAVP oral tablets is very low, between 0.08 and 0.16%. The bioavailability of oral administration is 5% of intranasal and 0.16% of intravenous [158]. Despite the low oral bioavailability, dDAVP exerts antidiuretic effects due to its high affinity to V2. Moreover, the concomitant food intake might have a substantial effect on dDAVP bioavailability [159]. In turn, nasal administration is burdened with high fluctuations of dosage, which can cause unexpected and dangerous side effects [160]. Despite poor bioavailability, oral delivery routes are still considered as highly demanded. Very recently, Kottke et al. [161] demonstrated the production of minitablets (smaller than 3 mm) containing the precise dosage of dDAVP, which rapidly disintegrates and provides immediate drug release. Zupančič et al. [162] showed that self-emulsifying drug delivery systems containing dDAVP are stable in vitro to glutathione and α-chymotrypsin degradation and are non-toxic. Thus, it might be considered as a novel, potential delivery strategy for oral dDAVP administration. Currently, the medical application of dDAVP includes the treatment of central diabetes insipidus, primary nocturnal enuresis and nocturia [163]. Recently, it was shown that low dose desmopressin Noqdirna®, as lyophilizate, is highly effective in treatments of nocturia due to idiopathic nocturnal polyuria in adults [164], same as MINIRIN® MELT 1995 [165]. At much higher doses, dDAVP is used to treat coagulation disorders, such as hemophilia A and von Willebrand’s disease (VWD). Moreover, this peptide is applied to prevent excessive bleeding during surgical procedures [166]. Noteworthy, dDAVP application is considered to be safe and is correlated with few adverse effects, such as headaches, diarrhea and potentially serious conditions of hyponatremia. The last condition is feasible in adult patients with renal failures. Thus, they are exposed to prolonged antidiuretic action of dDAVP, which is mostly excreted from the body in urine [158].
As mentioned above, dDAVP is a common hemostatic agent that facilitates clotting cascades in patients with VWD and hemophilia A. In the latter disease, which is manifested with an insufficient level of clotting factor VIII in human blood, the administration of dDAVP is beneficial only in patients with detectable levels of this endogenous factor [167]. In cases of inherited bleeding disorder VWD, which is classified into several types, dDAVP is theoretically useful in the treatment of type 1 VWD (patients with functionally normal von Willebrand’s factor—vWF). In practice, however, patients react differently to dDAVP administration, and such a treatment is routinely preceded with responsiveness challenge [168,169]. dDAVP is not recommended in type 2A (patients with qualitatively abnormal factors) and even contraindicated in 2B type due to the risk of transient thrombocytopenia after its administration. In patients with type 3, there is no response due to undetectable levels of vWF [170].
dDAVP is the significantly potent and selective agonist of G protein–coupled V2 membrane receptor in rats and in humans. In the latter case, it is also a strong agonist of the V1b receptor with an even lower Ki value than compared to the V2 receptor [121]. Nevertheless, V2 is mostly considered as the main pharmacological target for dDAVP. The functional extrarenal expression of V2 was also demonstrated in lung endothelial cells [171,172], as well as in many types of cancer cells, including lung, colorectal and breast cancer [173,174,175,176]. dDAVP, as a strong agonist of V2, has been repurposed as an adjuvant agent in the treatment of various tumors. It was reported to reduce tumor-induced angiogenesis in the breast cancer model [177]. The intravenous administration of dDAVP into Balb/c mice bearing aggressive F3II mammary carcinoma, resulted in an increased formation of angiostatin by tumor cells. The enhanced formation of angiostatin, which is a potent natural inhibitor of angiogenesis produced by cancer-mediated proteolysis of plasminogen, is likely associated with dDAVP-induced secretions of plasminogen activators such as uPA. As described by Gately et al. [178] such activators participate in the generation of angiostatin.
After the promising results of studies with mouse models [179,180] and two different veterinary trials in dogs with locally advanced mammary tumors [181,182], which demonstrated a significant reduction in metastatic progression and survival benefits, a phase II dose-escalation trial (NCT01606072) was conducted [183]. In that research, dDAVP (lyophilized formulation in saline) was administered intravenously to 20 patients with breast cancer before and after surgical resection. As a result, reduced intraoperative bleeding and increased plasma level of vWF was reported. Noteworthy, the growth of vWF secretion is beneficial for both hemostatic [184] and antimetastatic activity [185]. Additionally, as was detected by quantitative real-time reverse transcription-PCR assay for the expression of cytokeratin-19 mRNA in whole blood, a short postoperative drop in circulating tumor cells counts was observed. After one month, the level of cytokeratin-19 transcript returned to the baseline detected before surgery. The next phase I/II trial (NCT01623206) demonstrated that dDAVP may be considered as a promising hemostatic agent in rectal cancer patients with bleeding [186]. After a short, two-day treatment with dDAVP most patients showed at least a partial hemostatic response while about 60% had complete absence of bleeding symptoms at fourth day after the treatment started. Such an effect was observed at the last follow-up on day 14. Perfusion of rectal tumor was reduced in two-thirds of patients after dDAVP administration. The most prominent treatment-related severe adverse event was hyponatremia, associated in some cases with increased blood pressure.
Very recently, Sobol et al. [187] showed for the first time that dDAVP exerts antitumor in vivo activity on MG-63 human osteosarcoma xenografts in immunocompromised mice. Upon sustained intravenous administration, a significant reduction in tumor volume over time was reported. As compared to the control group, 34% tumor growth rate inhibition and a 25% reduction in resected tumor weight were observed. As examined by body weight and blood biochemical and hematological analyses, the antitumor action was accompanied by good tolerability and safety. Interestingly, the inhibition of phosphodiesterases, which are responsible for cAMP degradation, by rolipram enhanced the antiproliferative effects of dDAVP in osteosarcoma cells [166].
A combination treatment of dDAVP with the cytotoxic agent, docetaxel, resulted in a significant reduction in the proliferation of castrate-resistant prostate cancer (CRPC) cell line DU145 [188]. Moreover, dDAVP enhanced the anti-migratory effect of the drug and increased the inhibition of tumor growth in a xenograft model of prostate cancer (the mean tumor volume was about 41% lower as compared to treatment with the docetaxel alone). The effectiveness of that combined therapy was also confirmed using an in vivo orthotopic mouse model of CRPC [189].

Selepressin

Selepressin (55, Figure 8) [125] is another synthetic peptide analogue of AVP in which the following substitutions were applied Tyr2 → Phe2, Phe3 → Ile3, Gln4 → hGln4 and Arg8 → Orn(iPr)8 (Table 4). Selepressin may be used in the treatment of patients with septic shock during sepsis [125,127]. Sepsis is characterized by vasodilation and increased capillary permeability, which results in hypotension and loss of intravascular fluid and, subsequently, death [125,127]. The main causes of its mortality are low endothelial permeability and vascular hyperpermeability, which lead to organ dysfunction [126]. Thrombin, vascular endothelial growth factor (VEGF), angiopoietin 2 and LPS-induced pulmonary microvascular endothelial barrier disruptor cause endothelial dysfunction. Selepressin counteracts the effects of these disruptors and improves endothelial barrier function [126]. It also upregulates the expression of p53 (tumor suppressor gene), which is involved in the enhancement of the endothelial barrier. Selepressin is a potent vasopressor and it is selective for the V1a receptor [190,191]. The latter was demonstrated during phase I clinical trials in humans, when selepressin showed V1a-agonistic vasopressor properties and no signs of V2 activity [192,193]. In the phase IIa randomized, placebo-controlled trial in patients with septic shock, selepressin maintained adequate arterial pressure in the absence of norepinephrine and increased the proportion of patients that no longer require mechanical ventilation [127,193]. In phase IIb-III, selepressin was tested in vasopressor-dependent septic shock patients for its ability to reduce the duration of vasopressor and mechanical ventilator treatment. The administration of selepressin did not cause improvements compared with placebos, and the testing phase was terminated [192].
The present study indicates that the application of selepressin may be a potential alternative to AVP and serve as a supplementary vasopressor, especially to increase arterial pressure, prevent microvascular leak and reduce pulmonary edema [127]. However, further research is needed to confirm the role of selepressin.

Felypressin

Felypressin, also called octapressin, (56, Figure 8, Table 4) is known to have potent vasoconstricting properties and low toxicity. It works on smooth muscles [128], is able to constrict coronary vessels [194] and is an antidiuretic agent [129], although the last two activities are weaker than in the case of the parent peptide, AVP.
Because of its vasoconstricting properties, felypressin is used in anesthesia in dentistry and medicine. It promotes local vasoconstriction, thus enhancing the painkilling effect and decreasing bleeding during surgery [195,196,197]. It is used as a replacement of epinephrine and adrenaline due to fewer side effects [198,199] and higher median lethal dose (LD50) [200]. Inagawa et al. [130] studied the effect of epinephrine and felypressin on myocardial oxygen balance when applied in dental anesthesia at routine doses. It was found that felypressin is responsible for lowering myocardial tissue oxygen tension, heart rate and aortic and myocardial tissue blood flow. In the case of epinephrine, both aortic and myocardial tissue blood flows increased, while myocardial tissue oxygen tension remained unchanged.
Even though felypressin has a hemodynamic effect, the amount that is applied during dental procedures is too small to observe any disturbances [198] for hypertensive patients [201]. Prilocaine with felypressin is an anesthetic agent of choice even in case of very high-risk patients suffering multiple disorders [202]. Interestingly, when the influence of prilocaine/felypressin anesthetic on the autonomic nervous system during extraction of impacted mandibular third molar was compared to lidocaine/adrenaline, the differences in circulation dynamics were observed. It was explained by increased sympathetic nervous activity in case of the first and a decrease in parasympathetic nervous activity in case of the second local anesthetic agent [203]. Prilocaine/felypressin application is also recommended for adults over 65 years old in cases when tachycardia needs to be avoided [204]. Despite the general safety of felypressin, animal studies indicated myocardial ischemia [130] and, in some cases, reduced coronary blood flow [131] caused by this peptide.

Ornipressin

Ornipressin is an AVP analogue where Arg2 is replaced with Orn (57, Figure 8, Table 4). Similarly to other AVP analogues, it results in vasoconstriction and acts on microcirculation. The application of vasoconstrictors before laparoscopic myomectomy has already been recommended [136]. Ornipressin half-life is 1–2 h and, indeed, Assaf et al. [137] reported reduced blood loss during laparoscopic myomectomy if ornipressin was applied.
Similarly, during minimally invasive myomectomy, ornipressin was reported to significantly reduce blood loss and the need for transfusion and only oxytocin was observed to provide better results in the ranking score of treatments (p-score = 0.92 vs. p-score = 0.93, respectively) [138]. Additionally, ornipressin was the most successful in shortening hospital stay but only when combined with tranexamic acid (TXA). Interestingly, if applied alone, it was one of the least effective.
Vasoconstrictors, including ornipressin (similarly to terlipressin), proved to be useful also in liver cirrhosis if hepatorenal syndrome (HRS) was diagnosed [205]. HRS is caused primarily by systemic circulatory dysfunction, but cirrhotic cardiomyopathy and systemic inflammation also contribute to this renal failure [206].

Terlipressin

Terlipressin (glypressin, N-tryglycyl-8-lysine-vasopressin, 58, Figure 8, Table 4) is a pro-drug for the endogenous/natural porcine hormone LVP [87,132]. It is converted in the human body to its active metabolite, LVP, when three N-terminal glycines are cleaved by endothelial peptidases. Terlipressin is reported as a long-acting AVP analogue. Its plasma half-life is 240–360 min [64] and the compound reaches the highest concentration in plasma 60–120 min after administration. Drug metabolism is mediated by endo- and exopeptidases. Only about 1% of the drug is excreted unchanged by the kidneys [133,134]. The main mechanism of its action is to stimulate V1 receptors located predominantly in vascular smooth muscle cells within the splanchnic circulation, resulting in splanchnic vasoconstriction. The terlipressin-induced splanchnic vasoconstriction induces increased renal blood flow and has beneficial effects on HRS. It also reduces portal pressure and plays a role in reducing the risk of portal hypertensive bleeding [135,207].
Terlipressin also has a trace ability to stimulate V2 [208]. For this reason, drug-induced hyponatremia should be taken into account in patients receiving terlipressin [209,210,211]. There is also experimental evidence that the stimulation of V2 vasopressin receptors causes dilation of the arcuate vessels of the kidneys with an increase in blood flow in the parenchymal layer of the kidney [212]. The myoconstrictive effect of terlipressin affects not only the vascular muscle but also the smooth muscles outside the bloodstream. It results, inter alia, in an increase in the peristalsis of the esophagus, stomach and intestines, as well as in the intensification of uterine contractile activity [207].
Apart from the direct influence on V1 receptors, other indirect effects of terlipressin’s vasoconstricting action are also postulated. One of them is the inhibition of the production of nitric oxide by the endothelium [213]. Another one is the blocking of ATP-dependent potassium channels found in vascular smooth muscle during in vitro studies [214]. The visceral vasoconstriction and reduced blood flow to the portal vein reduce portal pressure, including a decrease in pressure in esophageal varices, while increasing total hepatic blood flow [215,216]. Terlipressin infusion reduces blood pressure in oesophageal varices by 35%, and the effect is greater when the portal pressure is higher [217,218,219]. Due to these properties, terlipressin is a drug recommended in the treatment of varicose bleeding in patients with liver cirrhosis [220,221,222]. Terlipressin is much better tolerated in liver cirrhosis than AVP; however, it may cause side effects, especially in the group of patients with cardiac burden. When using terlipressin, stenocardial complaints may be intensified. On the other hand, complications such as hypertensive crisis, dangerous cardiac arrhythmias and acute limb or intestinal ischemia are less frequent [221]. Patients treated with terlipressin may complain of abdominal pain, diarrhea, foot pain or bradycardia [222].
The unique pharmacological effects of terlipressin on circulation in cirrhosis include the reduction in portal pressure and increase in renal blood flow. These effects have been exploited in studies exploring its role in the management of HRS and variceal bleeding [207,223]. In recent years, as the physiological effects of terlipressin became better understood, roles of the drug in the setting of refractory ascites and cirrhotic hyponatremia have been proposed [224,225,226]. Despite the fact that terlipressin has been used for over 20 years, research on this drug is still continued. There are currently more than 50 ongoing clinical studies analyzing the further expansion of the role of this drug. For clinicians involved in the management of patients with advanced liver disease, terlipressin plays a central role in the management of complications and its participation is likely to expand in the coming years [135].
To overcome the rapid clearance of the polypeptide from blood circulation and its in vivo instability, Wang et al. [227] utilized chemical modifications of terlipressin with polyethylene glycol (PEG). The reaction was conducted in different pH value buffers at different molar ratios and PEGylation degree and position were determined. It turned out, that the highest amount of mono-PEG-terlipressin was obtained in the reaction between terlipressin and succinimidyl propionate-monmethoxy polyethylene glycol (m-PEG-SPA) performed in the lower content of PEG and in low pH value. Synthesis of di-PEG-terlipressin was the most effective at higher pH values and higher contents of PEG [227]. Furthermore, mono-PEG-terlipressin showed proteolytic cleavage during tryptic digestion, while di-PEG-terlipressin was resistant to it.

2.2. AVP and Its Analogues in Treatment of SARS-CoV-2

In 2019, a new severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) appeared and caused a new disease: COVID-19 [228]. The pathophysiological role of SARS-CoV-2 infection is mediated by abnormal immune response, endothelial dysfunction, direct viral toxicity and thrombo-inflammation, which lead to pulmonary manifestation [229]. Since the beginning of the pandemic, more attention has been focused on people with increased risk of morbidity and mortality caused by different comorbidities. Long-term exposure to SARS-CoV-2 infection leads to complex health problems in endocrine and cardiovascular systems [230]. Furthermore, it was shown that body functions regulated by AVP (regulation of the blood osmotic system, blood pressure, plasma volume and body water content) are disrupted during COVID-19 and related with poor clinical outcomes [229]. Therefore, the investigation of the relationship between AVP or its analogues and SARS-CoV-2 infection became high priority.
AVP was proved to be produced in COVID-19 patients in response to fever, dehydration, pain, physiological stress and high concentrations of pro-inflammatory cytokines, in order to oppose high blood viscosity [229]. High levels of AVP results in hyponatremia and inflammatory disorder. Its low concentrations, it leads to immunomodulatory effects, mainly in the lungs, through V2 receptor and further to the development of complications in COVID-19. AVP is present in the immune cells and can be released in response to inflammation and stress. AVP receptors are expressed on the immune cells involved in the release of pro-inflammatory cytokines and antibody production. They can be inhibited by vasopressin receptor antagonists (VRAs). These include the following: non-selective conivaptan (65, Figure 10) (blocks V1a and V2 receptors), V1a receptor antagonist relcovaptan (66, Figure 10), V1b receptor antagonist nelivaptan (67, Figure 10) and V2 receptor antagonist tolvaptan (68, Figure 10). In addition to the inhibition of AVP receptors, an in silico study showed that conivaptan inhibits SASRS-CoV-2 3C-like protease and viral RNA-dependent polymerase. In addition, it revealed that tolvaptan has anti-inflammatory and anti-fibrotic effects through the inhibition of monocyte chemotractic protein-1 and transforming growth factor β1, which are involved in the inflammatory process during SARS-CoV-2 infection. Therefore, AVP antagonists are considered as potential therapeutic agents for SARS-CoV-2 infection treatment.
Various in silico studies confirmed the potential of AVP analogues in COVID-19 treatment. Maffucci and Contini [231], in their recent work, used virtual screening (VS) to facilitate drug repurposing against SARS-CoV-2, targeting viral main proteinase and spike protein with 3000 existing drugs. The VS campaign on Spike protein Receptor Binding Domain (RBD) indicated that pressure regulators, i.e., terlipressin and lypressin, identified herein as potential binders of the S-protein may also be evaluated against SARS-CoV-2. Lypressin’s and ornipressin’s potential in the treatment of this intricate infection was confirmed. They were reported to have strong binding activity to both single chain core and the complex form (holoenzyme) of the SARS-CoV-2 RNA dependent RNA polymerase (RdRp) [232]. A docking study revealed a docking score of −11.717 kcal/mol for ornipressin and −11.923 kcal/mol for lypressin. Due to potent interaction with both forms of RdRp, the abovementioned AVP analogues are suggested to be tested as potential treatments of SARS-CoV-2 infection.
In addition, Sheikh et al. [230] reported a case of a patient who had a high serum sodium level and low urine osmolality, which were symptoms of diabetes insipidus developed after long-term exposure to the virus. Upon the administration of 2 µg of desmopressin, the patient improved clinically and symptomatically.

3. Conclusions

In this review, we presented the most relevant information regarding arginine vasopressin (AVP) and its analogues. AVP acts on three different G protein-coupled receptors (V1a, V1b and V2) and, depending on the interaction, demonstrates different functions. They include vasoconstriction, glycogenolysis, modulation of ACTH synthesis, stimulation of water reabsorption, insulin secretion, regulation of blood pressure, etc. Its expression in the brain acts on social interaction, depression and aggression. The first isolation, sequencing and chemical synthesis of AVP, conducted in the 1950s by du Vigneaud and his team, is deemed to be a landmark in organic chemistry. The subsequent identification of its receptors in the 1980s and 1990s gave rise to further exploration and understanding of biological properties displayed by this peptide hormone, as well as its natural (lypressin and phenypressin) and synthetic analogues (desmopressin, selepressin, felypressin, ornipressin and terlipressin) and many other synthetic non-peptide and peptide analogues.
Despite its extensive application in medicine, the use of AVP still has serious limitations, such as short biological half-life and lack of specificity for receptors resulting in side effects (e.g., hyponatremia, decrease of cardiac output and platelet count). In addition to AVP, its several efficient and relatively safe analogues were approved as medications. Felypressin is successfully used as anesthesia in medicine and dental care. Terlipressin is recommended in the treatment of varicose bleeding in patients with liver cirrhosis. dDAVP is utilized in the treatment of polyuric conditions including primary nocturnal enuresis, nocturia and diabetes insipidus as well as coagulation disorders, such as hemophilia A and von Willebrand’s disease. Compared to AVP, non-peptide analogue belovaptan showed amelioration of socialization and communication. Peptides lypressine (pigs) and phenypressin (marsupials) are characterized by easier applications in the treatment of some diseases, while having the same or better effect than AVP. Moreover, the interaction of analogues of AVP and LVP with receptors helped in the determination of key amino acids residues for their function and selectivity. It was shown that the activity and selectivity of AVP analogue responses to receptors depend on the steric size of group at the key positions (two, three, seven and eight) in peptides. The steric size of the side chain in position three of LVP plays significant role in the manifestation of vasopressin-like activities, but is not crucial for selectivity towards receptors. As in the case of many other known drugs, AVP and its analogues also were subjected to a repurposing strategy aimed at finding their novel applications. In this manner, desmopressin analogue [Val4,Gln5]dDAVP was found to possess not only antidiuretic but also anti-cancer activity. Lypressin, ornipressin and desmopressin showed satisfactory results in the treatment of SARS-CoV-2.
For several decades, exogenous AVP has been administered in the treatment of various diseases, including variceal bleeding, diabetes insipidus and vasodilatory shock. However, its non-specific binding to V1 and V2 receptors may promote potentially dangerous side effects. Even though many attractive AVP substitutes were developed and successfully applied in medicine, a quest for novel, highly specific AVP analogues remains an area of great interest. The search for AVP counterparts presenting either agonistic or antagonistic properties, characterized with proper ratio of structural complexity, activity and low toxicity, is still an open field.

Author Contributions

Conceptualization, A.G.-D. and M.G.; writing—original draft preparation, A.G, M.G., D.D., N.P. and A.Ł.; writing—review and editing, A.G.-D., D.D., M.G. and K.R.; visualization, A.G.-D. and M.G.; supervision, K.R. and A.Ł.; project administration, A.G.-D.; funding acquisition, A.G.-D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Science Center Poland (NCN), grant number UMO 2019/35/D/NZ7/00174 (A.G.-D.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Koopmanschap, G.; Ruijter, E.; Orru, R.V.A. Isocyanide-based multicomponent reactions towards cyclic constrained peptidomimetics. Beilstein J. Org. Chem. 2014, 10, 544–598. [Google Scholar] [CrossRef] [PubMed]
  2. Grauer, A.; König, B. Peptidomimetics—A versatile route to biologically active compounds. Eur. J. Org. Chem. 2009, 30, 5099–5111. [Google Scholar] [CrossRef]
  3. Vaisman, A.; Lushchak, O. Geroscience. In Encyclopedia of Biomedical Gerontology, Reference Modul in Biomedical Science, 1st ed.; Elsevier Inc. Academic Press: Aalborg, Denmark, 2020; pp. 154–159. [Google Scholar] [CrossRef]
  4. Dawgul, M.A.; Greber, K.E.; Sawicki, W.; Kamysz, W. Human host defense peptides—Role in maintaining human homeostasis and pathological processes. Curr. Med. Chem. 2017, 24, 654–672. [Google Scholar] [CrossRef]
  5. Lenci, E.; Trabocchi, A. Peptidomimetic toolbox for drug discovery. Chem. Soc. Rev. 2020, 49, 3262–3277. [Google Scholar] [CrossRef] [PubMed]
  6. Galicia-Garcia, U.; Benito-Vicente, A.; Jebari, S.; Larrea-Sebal, A.; Siddiqi, H.; Uribe, B.K.; Ostolaza, H.; Martín, C. Pathophysiology of type 2 diabetes mellitus. Int. J. Mol. Sci. 2020, 21, 6275. [Google Scholar] [CrossRef] [PubMed]
  7. Cantley, J.; Ashcroft, F.M. Q&A: Insulin secretion and type 2 diabetes: Why do β-cells fail? BMC Biol. 2015, 13, 33. [Google Scholar] [CrossRef] [Green Version]
  8. Sun, X.; Chen, W.D.; Wang, Y.-D. β-Amyloid: The key peptide in the pathogenesis of Alzheimer’s disease. Front. Pharmacol. 2015, 6, 221. [Google Scholar] [CrossRef] [Green Version]
  9. Chen, G.-F.; Xu, T.-H.; Zhou, Y.-R.; Jiang, Y.; Melcher, K.; Xu, H.E. Amyloid beta: Structure, biology and structure-based therapeutic development. Acta Pharmacol. Sin. 2017, 38, 1205–1235. [Google Scholar] [CrossRef]
  10. Martin, V.V.; Egelund, P.G.H.; Johansson, H.; Thordal Le Quement, S.; Wojcik, F.; Sejer Pedersen, D. Greening the synthesis of peptide therapeutics: An industrial perspective. RSC Adv. 2020, 10, 42457–42492. [Google Scholar] [CrossRef]
  11. Lee, A.C.-L.; Harris, J.L.; Khanna, K.K.; Hong, J.H. A comprehensive review on current advances in peptide drug development and design. Int. J. Mol. Sci. 2019, 20, 2383. [Google Scholar] [CrossRef] [Green Version]
  12. Lau, J.L.; Dunn, M.K. Therapeutic peptides: Historical perspectives, current development trends, and future directions. Bioorg. Med. Chem. 2018, 26, 2700–2707. [Google Scholar] [CrossRef] [PubMed]
  13. Adessi, C.; Soto, C. Converting a Peptide into a Drug: Strategies to Improve Stability and Bioavailability. Curr. Med. Chem. 2005, 9, 963–978. [Google Scholar] [CrossRef] [PubMed]
  14. Albericio, F.; Kruger, H.G. Therapeutic peptides. Future Med. Chem. 2012, 4, 1527–1531. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Henninot, A.; Collins, J.C.; Nuss, J.M. The Current State of Peptide Drug Discovery: Back to the Future? J. Med. Chem. 2018, 61, 1382–1414. [Google Scholar] [CrossRef] [PubMed]
  16. Räder, A.F.B.; Weinmüller, M.; Reichart, F.; Schumacher-Klinger, A.; Merzbach, S.; Gilon, C.; Hoffmann, A.; Kessler, H. Orally Active Peptides: Is There a Magic Bullet? Angew. Chemie-Int. Ed. 2018, 57, 14414–14438. [Google Scholar] [CrossRef]
  17. Chia, C.S.B. A Review on the Metabolism of 25 Peptide Drugs. Int. J. Pept. Res. Ther. 2021, 27, 1397–1418. [Google Scholar] [CrossRef]
  18. Hamman, J.H.; Enslin, G.M.; Kotzé, A.F. Oral delivery of peptide drugs: Barriers and developments. BioDrugs 2005, 19, 165–177. [Google Scholar] [CrossRef] [PubMed]
  19. Patel, A.; Patel, M.; Yang, X.; Mitra, A. Recent advances in protein and Peptide drug delivery: A special emphasis on polymeric nanoparticles. Protein Pept. Lett. 2014, 21, 1102–1120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Huttunen, K.M.; Raunio, H.; Rautio, J. Prodrugs-from serendipity to rational design. Pharmacol. Rev. 2011, 63, 750–771. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Yoshimura, M.; Conway-Campbell, B.; Ueta, Y. Arginine vasopressin: Direct and indirect action on metabolism. Peptides 2021, 142, 170555. [Google Scholar] [CrossRef]
  22. Garrahy, A.; Thompson, C.J. Vasopressin. In Encyclopedia of Endocrine Diseases, 2nd ed.; Academic Press: Oxford, UK, 2018; pp. 29–35. [Google Scholar] [CrossRef]
  23. Russell, J.A. Vasopressor therapy in critically ill patients with shock. Intensive Care Med. 2019, 45, 1503–1517. [Google Scholar] [CrossRef]
  24. Bankir, L.; Bichet, D.G.; Morgenthaler, N.G. Vasopressin: Physiology, assessment and osmosentation. J. Int. Med. 2017, 282, 284–297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Wei, H.-H.; Yuan, X.S.; Chen, Z.-K.; Chen, P.-P.; Xiang, Z.; Qu, W.-M.; Li, R.-X.; Zhou, G.-M.; Huang, Z.-H. Presynaptic inputs to vasopressin neurons in the hypothalamic supraoptic nucleus and paraventricular nucleus in mice. Exp. Neurol. 2021, 343, 113784. [Google Scholar] [CrossRef] [PubMed]
  26. Russell, J.A.; Gordon, A.C.; Williams, M.D.; Oyd, J.H.; Walley, K.R.; Kissoon, N. Vasopressor Therapy in the Intensive Care Unit. Semin. Respir. Crit. Care Med. 2021, 42, 59–77. [Google Scholar] [CrossRef] [PubMed]
  27. Mitra, A.K. Oxytocin and vasopressin: The social networking buttons of the body. AIMS Mol. Sci. 2021, 8, 32–50. [Google Scholar] [CrossRef]
  28. Oliver, H.; Shafer, E. On the physiological action of extracts of the pituitary body and certain other glandular organs. J. Physiol. 1895, 18, 277–279. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. du Vigneaud, V.; Popenoe, E.A.; Lawler, H. Partial purification and amino acid content of vasopressin from hog posterior. pituitary glands. J. Am. Chem. Soc. 1952, 74, 3713. [Google Scholar] [CrossRef]
  30. Berde, B.; Weidmann, H.; Cerletti, A. On phenylalanine-2-lysine-vasopressin. Helv. Physiol. Pharmacol. Acta 1961, 19, 285. [Google Scholar] [PubMed]
  31. Huguenin, R.L.; Boissonnas, R.A. Syntheses de la Phe2-arginine-vasopressine et de la Phe2-arginine vasotocineet nouvelles synthese de l’argininevasopressine et de l’arginine-vasotocine. Helv. Chim. Acta 1962, 45, 1629. [Google Scholar] [CrossRef]
  32. Berde, B.; Boissonnas, R.A.; Hueuenin, R.L.; Strümer, E. Vasopressin analogues with selective pressor activity. Experientia 1964, 20, 42–43. [Google Scholar] [CrossRef] [PubMed]
  33. Kasafire, E.; Rabek, V.; Rudinger, J.; Šorm, F. Amino acids and peptides. Synthesis of ten extended-chain analogues of lysine vasopressin. Coll. Czech. Chem. Commun. 1966, 31, 4581. [Google Scholar] [CrossRef]
  34. Chauvet, M.T.; Hurpet, D.; Chauvet, J.; Acher, R. Phenypressin (Phe2-Arg8-vasopressin), a new neurohypophysial peptide found in marsupials. Nature 1980, 287, 640–642. [Google Scholar] [CrossRef] [PubMed]
  35. Wisniewski, K.; Galyean, R.; Croston, G.; Laporte, R.; Taki, H.; Schteingart, C.D.; Rivière, P.-J.M. Synthesis and in vitro. pharmacological profile of potent and selective peptidic V1a receptor agonists. In Pept. Youth; Springer: New York, NY, USA, 2009; pp. 507–508. [Google Scholar] [CrossRef]
  36. Callahan, J.F.; Ashton-Shue, D.; Bryan, H.G.; Bryan, W.M.; Heckman, G.D.; Kinter, L.B.; McDonald, J.E.; Moore, M.L.; Schmidt, D.B.; Silvestri, J.S.; et al. Structure-Activity Relationships of Novel Vasopressin Antagonists Containing C-Terminal Diaminoalkanes and (Aminoalkyl)guanidines. J. Med. Chem. 1989, 32, 391–396. [Google Scholar] [CrossRef] [PubMed]
  37. Rubini, C.; Osler, A.; Calderan, A.; Guiotto, A.; Ruzza, P. Mechanistic studies of amide bond scission during acidolytic deprotection of Pip containing peptide. J. Pept. Sci. 2008, 14, 989–997. [Google Scholar] [CrossRef] [PubMed]
  38. Dekan, Z.; Kremsmayr, T.; Keov, P.; Godin, M.; Teakle, N.; Dürrauer, L.; Xiang, H.; Gharib, D.; Bergmayr, C.; Hellinger, R.; et al. Nature-inspired dimerization as a strategy to modulate neuropeptide pharmacology exemplified with vasopressin and oxytocin. Chem. Sci. 2021, 12, 4057–4062. [Google Scholar] [CrossRef] [PubMed]
  39. Wilds, A.L.; Ralls, W.; Tyner, A.; Daniel, R.; Kraychy, S.; Harnik, M. The synthesis of an octapeptide amide with the hormonal activity of oxytocin. J. Am. Chem. Soc. 1953, 75, 4879–4880. [Google Scholar]
  40. Alexander, R.; Aragón, O.R.; Bookwala, J.; Cherbuin, N.; Gatt, J.M.; Kahrilas, I.J.; Kästern, N.; Lawrence, A.; Lowe, L.; Morrison, R.G.; et al. The neuroscience of positive emotions and affect: Implications for cultivating happiness and wellbeing. Neurosci. Biobehav. Rev. 2021, 121, 220–249. [Google Scholar] [CrossRef] [PubMed]
  41. Goh, K.K.; Chen, C.-H.; Lane, H.-Y. Oxytocin in schizophrenia: Pathophysiology and implications for future treatment. Int. J. Mol. Sci. 2021, 22, 2146. [Google Scholar] [CrossRef] [PubMed]
  42. Gimpl, G.; Fahrenholz, F. The oxytocin receptor system: Structure, function, and regulation. Physiol. Rev. 2001, 81, 629–683. [Google Scholar] [CrossRef] [Green Version]
  43. Huang, Y.; Huang, X.; Ebstein, R.P.; Yu, R. Intranasal oxytocin in the treatment of autism spectrum disorders: A multilevel meta-analysis. Neurosci. Biobehav. Rev. 2021, 122, 18–27. [Google Scholar] [CrossRef]
  44. Land, H.; Schütz, G.; Schmale, H.; Richter, R. Nucleotide sequence of cloned cDNA encoding bovine arginine vasopressin-neurophysin II precursor. Nature 1982, 295, 299–303. [Google Scholar] [CrossRef] [PubMed]
  45. Küchler, S.; Perwitz, N.; Schick, R.R.; Klein, J.; Westphal, S. Arginine-vasopressin directly promotes a thermogenic and pro-inflammatory adipokine expression profile in brown adipocytes. Regul. Pept. 2010, 164, 126–132. [Google Scholar] [CrossRef] [PubMed]
  46. Mohan, S.; Moffett, R.C.; Thomas, K.G.; Irwin, N.; Flatt, P.R. Vasopressin receptors in islets enhance glucose tolerance, pancreatic beta-cell secretory function, proliferation and survival. Biochimie 2019, 158, 191–198. [Google Scholar] [CrossRef]
  47. Cotte, N.; Balestre, M.-N.; Phalipou, S.; Hibert, M.; Manning, M.; Barberis, C.; Mouillac, B. Identification of residues responsible for the selective binding of peptide antagonists and agonists in the V2 vasopressin receptor. J. Biol. Chem. 1998, 273, 29462–29468. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Shimada, Y.; Taniguchi, N.; Matsuhisa, A.; Akane, H.; Kawano, N.; Suzuki, T.; Tobe, T.; Kakefuda, A.; Yatsu, T.; Tahara, A.; et al. Synthesis and biological activity of novel 4,4-difluorobenzazepine derivatives as non-peptide antagonists of the arginine vasopressin V 1A receptor. Bioorg. Med. Chem. 2006, 14, 1827–1837. [Google Scholar] [CrossRef] [PubMed]
  49. Zeynalov, E.; Jones, S.M.; Elliott, J.P. Vasopressin and vasopressin receptors in brain edema. Vitam. Horm. 2020, 113, 291–312. [Google Scholar] [CrossRef] [PubMed]
  50. Tsukamoto, I.; Koshio, H.; Akamatsu, S.; Kuramochi, T.; Saitoh, C.; Yatsu, T.; Yanai-Inamura, H.; Kitada, C.; Yamamoto, E.; Sakamoto, S.; et al. Preparation of (4,4-difluoro-1,2,3,4-tetrahydro-5H-1-benzazepin-5-ylidene)acetamide derivatives as novel arginine vasopressin V2 receptor agonists. Bioorg. Med. Chem. 2008, 16, 9524–9535. [Google Scholar] [CrossRef]
  51. Sobolewski, D.; Prahl, A.; Derdowska, I.; Kwiatkowska, A.; Slaninova, J.; Lammek, B. Influence of Conformationally Constrained Amino Acids Replacing Positions 2 and 3 of Arginine Vasopressin (AVP) and Its Analogues on Their Pharmacological Properties. Protein Pept. Lett. 2017, 14, 213–217. [Google Scholar] [CrossRef] [PubMed]
  52. Demiselle, J.; Fage, N.; Radermacher, P.; Asfar, P. Vasopressin and its analogues in shock states: A review. Ann. Intensive Care 2020, 10, 9. [Google Scholar] [CrossRef] [PubMed]
  53. Corbani, M.; Trueba, M.; Stoev, S.; Murat, B.; Mion, J.; Boulay, V.; Guillon, G.; Manning, M. Design, synthesis, and pharmacological characterization of fluorescent peptides for imaging human V1b vasopressin or oxytocin receptors. J. Med. Chem. 2011, 54, 2864–2877. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Estrada, E.F.; Barra, V.; Caorsi, C.E.; Troncoso, S.; González, C.B.; Ruiz-Opazo, N. Identification of the V1 Vasopressin Receptor by Chemical Cross-Linking and Ligand Affinity Blotting. Biochemistry 1991, 30, 8611–8616. [Google Scholar] [CrossRef] [PubMed]
  55. Łempicka, E.; Derdowska, I.; Kowalczyk, W.; Dawidowska, O.; Prahl, A.; Janecki, M.; Jasiński, T.; Trzeciak, H.I.; Lammek, B. Analogues of arginine vasopressin modified in position 2 and 3 with conformationally constrained dipeptide fragments. J. Pept. Sci. 2005, 11, 91–96. [Google Scholar] [CrossRef] [PubMed]
  56. Kowalczyk, W.; Prahl, A.; Dawidowska, O.; Sobolewski, D.; Hardtrodt, B.; Lammek, B. The influence of 1-aminocyclopentane-1-carboxylic acid at position 2 or 3 of AVP and its analogues on their pharmacological properties. J. Pept. Sci. 2005, 11, 584–588. [Google Scholar] [CrossRef] [PubMed]
  57. Cheetham, T.; Baylis, P.H. Diabetes insipidus in children: Pathophysiology, diagnosis and management. Pediatr. Drugs 2002, 4, 785–796. [Google Scholar] [CrossRef] [PubMed]
  58. Nielsen, N.; Fohiaer, J.; Marples, D.; Kwon, T.-H.; Agre, P.; Knepper, M.A. Aquaporins in the kidney: From molecules to medicine. Physiol. Rev. 2002, 1, 205–244. [Google Scholar] [CrossRef] [PubMed]
  59. Bockenhauer, D.; Bichet, D. Pathophysiology, diagnosis and management of nephrogenic diabetes insipidus. Nat. Rev. Nephrol. 2015, 11, 576–588. [Google Scholar] [CrossRef] [PubMed]
  60. Yea, C.M.; Allan, C.E.; Ashworth, D.M.; Barnett, J.; Baxter, A.J.; Broadbridge, J.D.; Franklin, R.J.; Hampton, S.L.; Hudson, P.; Horton, J.A.; et al. New benzylureas as a novel series of potent, nonpeptidic vasopressin V2 receptor agonists. J. Med. Chem. 2008, 51, 8124–8134. [Google Scholar] [CrossRef] [PubMed]
  61. Tsukamoto, I.; Koshio, H.; Kuramochi, T.; Saitoh, C.; Yanai-Inamura, H.; Kitada-Nozawa, C.; Yamamoto, E.; Yatsu, T.; Shimada, Y.; Sakamoto, S.; et al. Synthesis and structure-activity relationships of amide derivatives of (4,4-difluoro-1,2,3,4-tetrahydro-5H-1-benzazepin-5-ylidene)acetic acid as selective arginine vasopressin V2 receptor agonists. Bioorg. Med. Chem. 2009, 17, 3130–3141. [Google Scholar] [CrossRef]
  62. Holz, G.G.; Kang, G.; Harbeck, M.; Roe, M.W.; Chepurny, O.G. Cell physiology of cAMP sensor. Epac. J. Physiol. 2006, 577, 5–15. [Google Scholar] [CrossRef] [PubMed]
  63. Chou, C.-L.; Christense, B.M.; Frische, S.; Vorum, H.; Desai, R.A.; Hoffert, J.D.; de Lanerolle, P.; Nielsen, S.; Knepper, M.A. Non-muscle myosin II and myosin light chain kinase are downstream targets for vasopressin signaling in the renal collecting duct. J. Biol. Chem. 2004, 279, 49026–49035. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Favory, R.; Salgado, D.R.; Vincent, J.-L. Investigational vasopressin receptor modulators in the pipeline. Expert Opin. Investig. Drugs 2009, 18, 1119–1131. [Google Scholar] [CrossRef] [PubMed]
  65. Felmet, K.; Carcillo, J.A. Neuroendocrine-Immune Mediator Coordination and Disarray in Critical Illness. In Pediatric Critical Care; Mosby Inc.: Maryland Heights, MO, USA, 2006; pp. 1462–1473. [Google Scholar] [CrossRef]
  66. Narayen, G.; Mandal, S. Vasopressin receptor antagonists and their role in clinical medicine. Indian J. Endocrinol. Metab. 2012, 16, 183–191. [Google Scholar] [CrossRef]
  67. Masuda, S.; Hattori, A.; Matsumoto, H.; Miyazawa, S.; Natori, Y.; Mizutani, S.; Tsujimoto, M. Involvement of the V2 receptor in vasopressin-stimulated translocation of placental leucine aminopeptidase/oxytocinase in renal cells. Eur. J. Biochem. 2003, 270, 1988–1994. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Tahara, A.; Tsukada, J.; Tomura, Y.; Wada, K.; Kusayama, T.; Ishii, N.; Tanaka, A. Effect of YM471, a nonpeptide AVP receptor antagonist, on human coronary artery smooth muscle cells. Peptides 2002, 23, 1809–1816. [Google Scholar] [CrossRef]
  69. Kim, J.W.; Kim, G.; Kim, T.W.; Han, W.; Kim, M.S.; Jeong, C.Y.; Park, D.H. Hemodynamic changes following accidental infiltration of a high dose of vasopressin. J. Int. Med. Res. 2020, 48, 0300060520959494. [Google Scholar] [CrossRef] [PubMed]
  70. Antoni, F.A. Vasopressin as a Stress Hormone. In Stress: Neuroendocrinology and Neurobiology: Handbook of Stressseries; Academic Press: Cambridge, MA, USA, 2017; Volume 2, pp. 97–108. [Google Scholar] [CrossRef]
  71. de Franchis, R. Somatostatin, somatostatin analogues and other vasoactive drugs in the treatment of bleeding oesophageal varices. Dig. Liver Dis. 2004, 36, 93–100. [Google Scholar] [CrossRef] [PubMed]
  72. Vasostrict Información Española De la Droga. Available online: https://www.drugs.com/mtm_esp/vasostrict.html (accessed on 11 December 2021).
  73. Vasopressin Price of 9 Brands/Trade Names|Medindia. Available online: https://www.medindia.net/drug-price/vasopressin.htm (accessed on 11 December 2021).
  74. Empressin®—Amomed. Available online: https://www.amomed.com/product/empressin-3/?lang=en&cn-reloaded=1 (accessed on 11 December 2021).
  75. Reverpleg®—Amomed. Available online: https://www.amomed.com/product/empressin-3-2/?lang=fr (accessed on 11 December 2021).
  76. Xiao, X.; Zhu, Y.; Zhen, D.; Chen, X.M.; Yue, W.; Liu, L.; Li, T. Beneficial and side effects of arginine vasopressin and terlipressin for septic shock. J. Surg. Res. 2015, 195, 568–579. [Google Scholar] [CrossRef] [PubMed]
  77. Kowalczyk, W.; Sobolewski, D.; Prahl, A.; Derdowska, I.; Borovičková, L.; Slaninová, J.; Lammek, B. The effects of N-terminal part modification of arginine vasopressin analogues with 2-aminoindane-2-carboxylic acid: A highly potent V2 agonist. J. Med. Chem. 2007, 50, 2926–2929. [Google Scholar] [CrossRef] [PubMed]
  78. Kakefuda, A.; Suzuki, T.; Tobe, T.; Tsukada, J.; Tsukamoto, S. Synthesis and pharmacological evaluation of 5-(4-biphenyl)-3-methyl-4-phenyl-1,2,4-triazole derivatives as a novel class of selective antagonists for the human vasopressin V1a receptor. J. Med. Chem. 2002, 45, 2589–2598. [Google Scholar] [CrossRef] [PubMed]
  79. Guillon, C.D.; Koppel, G.A.; Brownstein, M.J.; Chaney, M.O.; Ferris, C.F.; Lu, S.; Fabio, K.M.; Miller, M.J.; Heindel, N.D.; Hunden, D.C.; et al. Azetidinones as vasopressin V1a antagonists. Bioorganic Med. Chem. 2007, 15, 2054–2080. [Google Scholar] [CrossRef] [Green Version]
  80. Schnider, P.; Bissantz, C.; Bruns, A.; Dolente, C.; Goetschi, E.; Jakob-Roetne, R.; Künnecke, B.; Mueggler, T.; Muster, W.; Parrott, N.; et al. Discovery of Balovaptan, a Vasopressin 1a Receptor Antagonist for the Treatment of Autism Spectrum Disorder. J. Med. Chem. 2020, 63, 1511–1525. [Google Scholar] [CrossRef] [PubMed]
  81. Search of: Balovaptan—List Results—ClinicalTrials.gov. Available online: https://clinicaltrials.gov/ct2/results?cond=&term=Balovaptan&cntry=&state=&city=&dist= (accessed on 11 December 2021).
  82. Kondo, K.; Ogawa, H.; Shinohara, T.; Kurimura, M.; Tanda, Y.; Kan, K.; Yamashita, H.; Nakamura, S.; Hirano, T.; Yamamura, Y.; et al. Novel design of nonpeptide AVP V2 receptor agonists: Structural requirements for an agonist having 1-(4-aminobenzoyl)-2,3,4,5-tetrahydro-1H-1-benzazepine as a template. J. Med. Chem. 2000, 43, 4388–4397. [Google Scholar] [CrossRef] [PubMed]
  83. Cho, H.; Murakami, K.; Nakanishi, H.; Fujisawa, A.; Isoshima, H.; Niwa, M.; Hayakawa, H.; Hase, Y.; Uchida, I.; Watanabe, H.; et al. Synthesis and Structure—Activity Relationships of 5, 6, 7, 8-Tetrahydro-4 H-thieno [3,2-b] azepine Derivatives: Novel Arginine Vasopressin Antagonists. J. Med. Chem. 2004, 47, 101–109. [Google Scholar] [CrossRef]
  84. Poulani, D. Vasopressin and Oxytocin, 1st ed.; Elsevier Science: Amsterdam, The Netherlands, 2002. [Google Scholar]
  85. Maxfield, F.R.; Scheraga, H.A. A Raman Spectroscopic Investigation of the Disulfide Conformation in Oxytocin and Lysine Vasopressin. Biochemistry 1977, 16, 4443–4449. [Google Scholar] [CrossRef] [PubMed]
  86. Lypressin. Available online: https://pubchem.ncbi.nlm.nih.gov/compound/Lypressin (accessed on 15 December 2021).
  87. Sharman, A.; Low, J. Vasopressin and its role in critical care. Contin. Educ. Anaesth. Crit. Care Pain 2008, 8, 134–137. [Google Scholar] [CrossRef] [Green Version]
  88. Dakters, J.G.; Yashon, D.; Purnell, E.W.; Volk, M.; White, R.J. Ultrasonography for Diagnosis of Orbital Tumors. JAMA J. Am. Med. Assoc. 1968, 203, 803–805. [Google Scholar] [CrossRef]
  89. Dingman, J.H.; Hauger-Klevene, J.F. Treatment of Diabetes Insipidus: Synthetic Lysine Vasopressin Nasal Solution. J. Clin. Endocr. 1964, 24, 550–553. [Google Scholar] [CrossRef] [PubMed]
  90. Acher, R. The nonmammalian-mammalian transition through neurohypophysial peptides. Peptides 1985, 6, 309–314. [Google Scholar] [CrossRef]
  91. Hurpet, D.; Chauvet, M.T.; Chauvet, J.; Acher, R. Marsupial hypothalamo-neurohypophyseal hormones. Int. J. Pept. Protein Res. 1982, 19, 366–371. [Google Scholar] [CrossRef]
  92. Jarial, K.D.S.; Bhansali, A.; Gupta, V.; Singh, P.; Mukherjee, K.K.; Sharma, A.; Vashishtha, R.K.; Sukumar, S.P.; Sachdeva, N.; Wlia, R. Diagnostic accuracy and comparison of BIPSS in response to lysine vasopressin and hCRH. Endocr. Connect. 2018, 7, 425–432. [Google Scholar] [CrossRef] [Green Version]
  93. Wang, H.; Ba, Y.; Xing, Q.; Cai, R.C. Differential diagnostic value of bilateral inferior Petrosal sinus sampling (BIPSS) in ACTH-dependent Cushing syndrome: A systematic review and Meta-analysis. BMC Endocr. Disord. 2020, 20, 143. [Google Scholar] [CrossRef] [PubMed]
  94. Stahl, G.L.; Smith, C.W.; Walter, R. Oxytocin and Lysine-vasopressin with N5,N5-Dialkylglutamine in the 4 Position: Effect of Introducing Sterically Hindered Groups into the Hydrophilic Cluster of Neurohypophyseal Hormones. J. Med. Chem. 1980, 23, 213–217. [Google Scholar] [CrossRef] [PubMed]
  95. Dyckes, D.F.; Ferger, M.F.; du Vigneaud, V.; Chan, W.Y. Synthesis and Some of the Pharmacological Properties of [4-Leucine]-8-lysine-vasopressin and [1-Deamino,4-leucine]-8-lysine-vasopressin. J. Med. Chem. 1973, 16, 843–847. [Google Scholar] [CrossRef]
  96. Gillessen, D.; du Vigneaud, V. The Synthesis and Pharmacological Properties of 4-Decarboxamido-8-lysine-vasopressin, 5-Decarboxamido-8-lysine-vasopressin, and Their 1-Deamino Analogues. J. Biol. Chem. 1967, 242, 4806–4812. [Google Scholar] [CrossRef]
  97. Pena, A.; Murat, B.; Trueba, M.; Ventura, M.A.; Wo, N.C.; Szeto, H.H.; Cheng, L.L.; Stoev, S.; Guillon, G.; Manning, M. Design and Synthesis of the First Selective Agonists for the Rat Vasopressin V1bReceptor: Based on Modifications of Deamino-[Cys]arginine Vasopressin at Positions 4 and 8. J. Med. Chem. 2007, 50, 835–847. [Google Scholar] [CrossRef] [PubMed]
  98. Lundell, E.O.; Ferger, M.F. Synthesis and some pharmacological properties of [1-beta-mercaptopropionic acid, 2-3,5-dibromo-L-Tyrosine)]-8-Lysine-Vasopressin. Bioorg. Chem. 1975, 4, 377–384. [Google Scholar] [CrossRef]
  99. Siedel, W.; Sturm, K.; Geiger, R. Syntheseeines vasopressorisch wirkenden Peptids: 0-Methyl-tyrosin2-Lysin8-Vasopressin (OMTLV). Eur. J. Inorg. Chem. 1963, 96, 1436–1440. [Google Scholar] [CrossRef]
  100. Larsson, L.E.; Lindeberg, G.; Melin, P.; Pliska, V. Synthesis of O-alkylated lysine-vasopressin, inhibitors of the antidiuretic response to lysine-vasopressin. J. Med. Chem. 1978, 21, 352–356. [Google Scholar] [CrossRef]
  101. Smith, C.W.; Ferger, M.F.; Chan, W.Y. Synthesis and some pharmacological properties of [3-.beta.-(2-thienyl)-L-alanine]-8-lysine-vasopressin. J. Med. Chem. 1975, 18, 822–825. [Google Scholar] [CrossRef] [PubMed]
  102. Banerjee, S.N.; Diamond, L.; Ressler, C.; Sawyer, W.H. [3-(1,4-Cyclohexadienyl)-L-alanine,8-lysine]vasopressin synthesis and some pharmacological properties. J. Med. Chem. 1979, 22, 1487–1492. [Google Scholar] [CrossRef]
  103. Smith, C.W.; Walter, R.; Stavropoulos, G.; Theodoropoulos, D. [5-(N4,N4-Dimethylasparagine),8-lysine]vasopressin: The first 5-position neurohypophyseal hormone analog to retain significant antidiuretic potency. J. Med. Chem. 1980, 23, 217–219. [Google Scholar] [CrossRef] [PubMed]
  104. Kordower, J.H.; Sikorszky, V.; Bodnar, R.J. Central antinociceptive effects of lysine-vasopressin and an analogue. Peptides 1982, 3, 613–617. [Google Scholar] [CrossRef]
  105. Glass, J.D.; Du Vigneaud, V. Solid-phase synthesis and pressor potency of [1-deamino-9-ethylenediamine]-lysine-vasopressin. J. Med. Chem. 1973, 16, 160–161. [Google Scholar] [CrossRef] [PubMed]
  106. Chauvet, M.T.; Hurpet, D.; Chauvet, J.; Acher, R. Identification of mesotocin, lysine vasopressin, and phenypressin in the eastern gray kangaroo (Macropus giganteus). Gen. Comp. Endocrinol. 1983, 49, 63–72. [Google Scholar] [CrossRef]
  107. Berde, B.; Boissonnas, R.A. Basic Pharmacological Properties of Synthetic Analogues and Homologues of the Neurohypophysial Hormones. In Neurohypophysial Hormones and Similar Polypeptides; Springer: Berlin/Heidelberg, Germany, 1968; pp. 802–870. [Google Scholar] [CrossRef]
  108. Kowalczyk, W.; Prahl, A.; Derdowska, I.; Dawidowska, O.; Slaninová, J.; Lammek, B. Highly Potent 1-Aminocyclohexane-1-Carboxylic Acid Substituted V2 Agonists of Arginine Vasopressin. J. Med. Chem. 2004, 47, 6020–6024. [Google Scholar] [CrossRef] [PubMed]
  109. Bankowski, K.; Manning, M.; Haldar, J.; Sawyer, W.H. Design of potent antagonists of the vasopressor response to arginine-vasopressin. J. Med. Chem. 1978, 21, 850–853. [Google Scholar] [CrossRef] [PubMed]
  110. Jones, D.A., Jr.; Sawyer, W.H. N-Acetyl-[2-(O-methyl)tyrosine]arginine-vasopressin, an interesting antagonist of the vasopressor response to vasopressin. J. Med. Chem. 1980, 23, 696–698. [Google Scholar] [CrossRef]
  111. Kwiatkowska, A.; Sobolewski, D.; Prahl, A.; Borovicková, L.; Slaninová, J.; Lammek, B. Arginine vasopressin and its analogues--the influence of position 2 modification with 3,3-diphenylalanine enantiomers. Highly potent V2 agonists. Eur. J. Med. Chem. 2009, 44, 2862–2867. [Google Scholar] [CrossRef]
  112. Lebl, M.; Brtnik, F. Tables of Analogs in: Handbook of Neurohypophyseal Hormone Analogs; Jošt, K., Lebl, M., Brtnik, F., Eds.; CRC Press: Boca Raton, FL, USA, 1987; Part 2; Volume 2, pp. 127–167. [Google Scholar]
  113. Lammek, B.; Czaja, M.; Derdowska, I.; Rekowski, P.; Trzeciak, H.; Sikora, P.; Szkróbka, W.; Stojko, R.; Kupryszewski, G. Influence of L-naphthylalanine in position 3 of AVP and its analogues on their pharmacological properties. J. Pept. Res. 1997, 49, 261–268. [Google Scholar] [CrossRef]
  114. Grzonka, Z.; Lammek, B.; Kasprzykowski, F.; Gazis, D.; Schwartz, I.L. Synthesis and some pharmacological properties of oxytocin and vasopressin analogues with sarcosine or N-methyl-L-alanine in position 7. J. Med. Chem. 1983, 26, 555–559. [Google Scholar] [CrossRef]
  115. Ali, F.E.F.; Bryan, W.; Chang, H.L.; Huffman, W.F.; Moore, M.L.; Heckman, G.; Kinter, L.B.; McDonald, J.; Schmidt, D. Potent vasopressin antagonists lacking the proline residue at position 7. J. Med. Chem. 1986, 29, 984–988. [Google Scholar] [CrossRef]
  116. Fink, M.L.; Bodanszyky, M. Synthesis and hormonal activities of 8-L-homonorleucine-vasopressin. J. Med. Chem. 1973, 16, 1324–1326. [Google Scholar] [CrossRef]
  117. Katsoyannis, P.G.; du Vigneaud, V. The synthesis of the histidine analog of the vasopressins. Arch. Biochem. Biophys. 1958, 78, 555–562. [Google Scholar] [CrossRef]
  118. Manning, M.; Coy, E.J.; Sawyer, W.H.; Acosta, M. Solid-phase synthesis and some pharmacological properties of 4-threonine analogs of vasopressins and vasotocin and of arginine-vasopressin and arginine-vasotocin. J. Med. Chem. 1973, 16, 463–466. [Google Scholar] [CrossRef]
  119. Paranjape, S.B.; Thibonnier, M. Development and therapeutic indications of orally-active non-peptide vasopressin receptor antagonists. Expert Opin. Investig. Drugs 2001, 10, 825–834. [Google Scholar] [CrossRef]
  120. Wang, J.; Yadav, V.; Smart, A.L.; Tajiri, S.; Basit, A.W. Toward Oral Delivery of Biopharmaceuticals: An Assessment of the Gastrointestinal Stability of 17 Peptide Drugs. Mol. Pharm. 2015, 12, 966–973. [Google Scholar] [CrossRef] [Green Version]
  121. Manning, M.; Stoev, S.; Chini, B.; Durroux, T.; Mouillac, B.; Guillon, G. Peptide and non-peptide agonists and antagonists for the vasopressin and oxytocin V1a, V1b, V2 and OT receptors: Research tools and potential therapeutic agents. Prog. Brain Res. 2008, 170, 473–512. [Google Scholar] [CrossRef]
  122. Zaoral, M.; Kolc, J.; Šorm, F. Amino acids and peptides. LXXI. Synthesis of 1-deamino-8-D-γ-aminobutyrine-vasopressin, 1-deamino-8-D-lysine-vasopressin, and 1-deamino-8-D-arginine-vasopressin. Collect. Czechoslov. Chem. Commun. 1967, 32, 1250–1257. [Google Scholar] [CrossRef]
  123. Krchnak, V.; Zaoral, M. Synthesis of (1-β-mercaptopropionic acid, 8-D-arginine) vasopressin (DDAVP) in solid phase. Simple optimization. Coll. Czech. Chem. Commun. 1979, 44, 1173–1178. [Google Scholar] [CrossRef]
  124. Manning, M. Impact of the Merrifield solid phase method on the design and synthesis of selective agonists and antagonists of oxytocin and vasopressin: A historical perspective. Biopolymers 2008, 90, 203–212. [Google Scholar] [CrossRef] [Green Version]
  125. Lewis, R.J.; Angus, D.C.; Laterre, P.-F.; Kjølbye, A.L.; Van Der Meulen, E.; Blemings, A.; Graves, T.; Russell, J.A.; Carlsen, J.E.; Jacobsen, K.; et al. Rationale and Design of an Adaptive Phase 2b/3 Clinical Trial of Selepressin for Adults in Septic Shock. Selepressin Evaluation Programme for Sepsis-induced Shock—Adaptive Clinical Trial. Ann. Am. Thorac. Soc. 2018, 15, 250–257. [Google Scholar] [CrossRef]
  126. Barabutis, N.; Marinova, M.; Solopov, P.; Uddin, M.A.; Croston, G.E.; Reinheimer, T.M.; Catravas, J.D. Protective Mechanism of the Selective Vasopressin V1A Receptor Agonist Selepressin against Endothelial Barrier Dysfunction. J. Pharmacol. Exp. Ther. 2020, 375, 286–295. [Google Scholar] [CrossRef]
  127. Milano, S.P.; Boucheix, O.B.; Reinheimer, T.M. Selepressin, a novel selective V1A receptor agonist: Effect on mesenteric flow and gastric mucosa perfusion in the endotoxemic rabbit. Peptides 2020, 129, 170318. [Google Scholar] [CrossRef]
  128. Berde, B.; Doepfner, W.; Konzett, H. Some Pharmacological Actions of Four Synthetic Analogues Of Oxytocin. Br. J. Pharmacol. Chemother. 1957, 12, 209–214. [Google Scholar] [CrossRef] [Green Version]
  129. Figallo, M.A.S.; Cayón, R.T.V.; Lagares, D.T.; Flores, J.R.C.; Portillo, G.M. Use of anesthetics associated to vasoconstrictors for dentistry in patients with cardiopathies. Review of the literature published in the last decade. J. Clin. Exp. Dent. 2012, 4, e107–e111. [Google Scholar] [CrossRef]
  130. Inagawa, M.; Ichinohe, T.; Kaneko, Y. Felypressin, but Not Epinephrine, Reduces Myocardial Oxygen Tension After an Injection of Dental Local Anesthetic Solution at Routine Doses. J. Oral Maxillofac. Surg. 2010, 68, 1013–1017. [Google Scholar] [CrossRef] [Green Version]
  131. Agata, H.; Ichinohe, T.; Kaneko, Y. Felypressin-induced reduction in coronary blood flow and myocardial tissue oxygen tension during anesthesia in dogs. Can. J. Anaesth. 1999, 46, 1070–1075. [Google Scholar] [CrossRef] [Green Version]
  132. Jamil, K.; Pappas, S.C.; Devarakonda, K.R. In vitro binding and receptor-mediated activity of terlipressin at vasopressin receptors V1 and V2. J. Exp. Pharmacol. 2018, 10, 1–7. [Google Scholar] [CrossRef] [Green Version]
  133. Nilsson, G.; Lindblom, P.; Ohlin, M.; Berling, R.; Vernersson, E. Pharmacokinetics of terlipressin after single i.v. doses to healthy volunteers. Drugs Under Exp. Clin. Res. 1990, 16, 307–314. [Google Scholar]
  134. Saner, F.H.; Canbay, A.; Gerken, G.; Broelsch, C.E. Pharmacology, clinical efficacy and safety of terlipressin in esophageal varices bleeding, septic shock and hepatorenal syndrome. Expert Rev. Gastroenterol. Hepatol. 2007, 1, 207–217. [Google Scholar] [CrossRef]
  135. Papaluca, T.; Gow, P. Terlipressin: Current and emerging indications in chronic liver disease. J. Gastroenterol. Hepatol. 2018, 33, 591–598. [Google Scholar] [CrossRef]
  136. Chang, F.-H.; Soong, Y.-K.; Cheng, P.-J.; Lee, C.-L.; Lai, Y.-M.; Wang, H.-S.; Chou, H.-H. Surgery: Laparoscopic myomectomy of large symptomatic leiomyoma using airlift gasless laparoscopy: A preliminary report. Hum. Reprod. 1996, 11, 1427–1432. [Google Scholar] [CrossRef] [Green Version]
  137. Assaf, A. Adhesions after laparoscopic myomectomy: Effect of the technique used. Gynaecol. Endosc. 2001, 8, 225–229. [Google Scholar] [CrossRef]
  138. Samy, A.; Raslan, A.N.; Talaat, B.; El Lithy, A.; El Sharkawy, M.; Sharaf, M.F.; Hussein, A.H.; Amin, A.H.; Ibrahim, A.M.; Elsherbiny, W.S.; et al. Perioperative nonhormonal pharmacological interventions for bleeding reduction during open and minimally invasive myomectomy: A systematic review and network meta-analysis. Fertil. Steril. 2020, 113, 224–233.e6. [Google Scholar] [CrossRef]
  139. Pastrian, M.B.; Guzmán, F.; Garona, J.; Pifano, M.; Ripoll, G.V.; Cascone, O.; Ciccia, G.N.; Albericio, F.; Gomez, D.E.; Alonso, D.F.; et al. Structure-activity relationship of 1-desamino-8-D-arginine vasopressin as an antiproliferative agent on human vasopressin V2 receptor-expressing cancer cells. Mol. Med. Rep. 2014, 9, 2568–2572. [Google Scholar] [CrossRef]
  140. Czaplewski, C.; Kaźmierkiewicz, R.; Ciarkowski, J. Molecular modeling of the human vasopressin V2 receptor/agonist complex. J. Comput. Aided. Mol. Des. 1998, 12, 275–287. [Google Scholar] [CrossRef]
  141. Kowalczyk, W.; Prahl, A.; Derdowska, I.; Sobolewski, D.; Olejnik, J.; Zabrocki, J.; Borovicková, L.; Slaninová, J.; Lammek, B. Analogues of Neurohypophyseal Hormones, Oxytocin and Arginine Vasopressin, Conformationally Restricted in the N-Terminal Part of the Molecule. J. Med. Chem. 2006, 49, 2016–2021. [Google Scholar] [CrossRef]
  142. Manning, M.; Misicka, A.; Olma, A.; Bankowski, K.; Stoev, S.; Chini, B.; Durroux, T.; Mouillac, B.; Corbani, M.; Guillon, G. Oxytocin and Vasopressin Agonists and Antagonists as Research Tools and Potential Therapeutics. J. Neuroendocrinol. 2012, 24, 609–628. [Google Scholar] [CrossRef] [Green Version]
  143. Sawyer, W.H.; Acosta, M.; Balaspiri, L.; Judd, J.; Manning, M. Structural Changes in the Arginine Vasopressin Molecule that Enhance Antidiuretic Activity and Specificity1. Endocrinology 1974, 94, 1106–1115. [Google Scholar] [CrossRef]
  144. Erdélyi, L.S.; Balla, A.; Patócs, A.; Tóth, M.; Várnai, P.; Hunyady, L. Altered Agonist Sensitivity of a Mutant V2 Receptor Suggests a Novel Therapeutic Strategy for Nephrogenic Diabetes Insipidus. Mol. Endocrinol. 2014, 28, 634–643. [Google Scholar] [CrossRef] [Green Version]
  145. Garona, J.; Pifano, M.; Ripoll, G.; Alonso, D.F.; Eiden, L.E. Development and therapeutic potential of vasopressin synthetic analog [V4Q5]dDAVP as a novel anticancer agent. Vitam. Horm. 2020, 113, 259–289. [Google Scholar] [CrossRef]
  146. Iannucci, N.B.; Ripoll, G.V.; Garona, J.; Cascone, O.; Ciccia, G.N.; Gomez, D.E.; Alonso, D.F. “Antiproliferative effect of 1-deamino-8-d-arginine vasopressin analogs on human breast cancer cells. Fut. Med. Chem. 2011, 3, 1987–1993. [Google Scholar] [CrossRef]
  147. Garona, J.; Pifano, M.; Pastrian, M.B.; Gomez, D.E.; Ripoll, G.V.; Alonso, D.F. Addition of vasopressin synthetic analogue [V4Q5]dDAVP to standard chemotherapy enhances tumour growth inhibition and impairs metastatic spread in aggressive breast tumour models. Clin. Exp. Metastasis 2016, 33, 589–600. [Google Scholar] [CrossRef] [Green Version]
  148. Garona, J.; Pifano, M.; Orlando, U.D.; Pastrian, M.B.; Iannucci, N.B.; Ortega, H.H.; Podesta, E.J.; Gomez, D.E.; Ripoll, G.V.; Alonso, D.F. The novel desmopressin analogue [V4Q5]dDAVP inhibits angiogenesis, tumour growth and metastases in vasopressin type 2 receptor-expressing breast cancer models. Int. J. Oncol. 2015, 46, 2335–2345. [Google Scholar] [CrossRef]
  149. Pifano, M.; Garona, J.; Capobianco, C.S.; Gonzalez, N.; Alonso, D.F.; Ripoll, G.V. Peptide Agonists of Vasopressin V2 Receptor Reduce Expression of Neuroendocrine Markers and Tumor Growth in Human Lung and Prostate Tumor Cells. Front. Oncol. 2017, 7, 11. [Google Scholar] [CrossRef] [Green Version]
  150. Garona, J.; Sobol, N.T.; Pifano, M.; Segatori, V.I.; Gomez, D.E.; Ripoll, G.V.; Alonso, D.F. Preclinical Efficacy of [V4 Q5 ]dDAVP, a Second Generation Vasopressin Analog, on Metastatic Spread and Tumor-Associated Angiogenesis in Colorectal Cancer. Cancer Res. Treat. 2019, 51, 438–450. [Google Scholar] [CrossRef] [Green Version]
  151. Wisniewski, K.; Qi, S.; Kraus, J.; Ly, B.; Srinivasan, K.; Tariga, H.; Croston, G.; La, E.; Wisniewska, H.; Ortiz, C.; et al. Discovery of Potent, Selective, and Short-Acting Peptidic V2 Receptor Agonists. J. Med. Chem. 2019, 62, 4991–5005. [Google Scholar] [CrossRef]
  152. Gasthuys, E.; Dossche, L.; Michelet, R.; Nørgaard, J.P.; Devreese, M.; Croubels, S.; Vermeulen, A.; Van Bocxlaer, J.; Walle, J.V. Pediatric Pharmacology of Desmopressin in Children with Enuresis: A Comprehensive Review. Pediatr. Drugs 2020, 22, 369–383. [Google Scholar] [CrossRef]
  153. Robson, W.; Leung, A.; Norgaard, J. The Comparative Safety of Oral Versus Intranasal Desmopressin for the Treatment of Children with Nocturnal Enuresis. J. Urol. 2007, 178, 24–30. [Google Scholar] [CrossRef]
  154. Chanson, P.; Salenave, S. Treatment of neurogenic diabetes insipidus. Ann. d’Endocrinologie 2011, 72, 496–499. [Google Scholar] [CrossRef]
  155. Juul, K.V.; Van Herzeele, C.; De Bruyne, P.; Goble, S.; Walle, J.V.; Nørgaard, J.P. Desmopressin melt improves response and compliance compared with tablet in treatment of primary monosymptomatic nocturnal enuresis. Eur. J. Pediatr. 2013, 172, 1235–1242. [Google Scholar] [CrossRef] [Green Version]
  156. Costantini, E.; Mearini, L.; Lazzeri, M.; Bini, V.; Nunzi, E.; di Biase, M.; Porena, M. Laparoscopic Versus Abdominal Sacrocolpopexy: A Randomized, Controlled Trial. J. Urol. 2016, 196, 159–165. [Google Scholar] [CrossRef] [PubMed]
  157. Lottmann, H.; Froeling, F.; Alloussi, S.; El-Radhi, A.S.; Rittig, S.; Riis, A.; Persson, B.-E. A randomised comparison of oral desmopressin lyophilisate (MELT) and tablet formulations in children and adolescents with primary nocturnal enuresis. Int. J. Clin. Pract. 2007, 61, 1454–1460. [Google Scholar] [CrossRef] [PubMed]
  158. Manufactured for: Sanofi-Aventis U.S.; LLC Bridgewater: New York, NY, USA, Patent Nos. 5,498,598; 5,500,413; 5,596,078; 5,674,850; 5,763,407. 2007. Available online: https://www.accessdata.fda.gov/drugsatfda_docs/label/2007/017922s038,018938s027,019955s013lbl.pdf (accessed on 20 October 2021).
  159. Rittig, S.; Jensen, A.R.; Jensen, K.T.; Pedersen, E.B. Effect of food intake on the pharmacokinetics and antidiuretic activity of oral desmopressin (DDAVP) in hydrated normal subjects. Clin. Endocrinol. 1998, 48, 235. [Google Scholar] [CrossRef]
  160. van Kerrebroeck, P.; Nørgaard, J.P. Desmopressin for the treatment of primary nocturnal enuresis. Pediatr. Health 2009, 3, 311–327. [Google Scholar] [CrossRef]
  161. Kottke, D.; Burckhardt, B.B.; Knaab, T.C.; Breitkreutz, J.; Fischer, B. Development and evaluation of a composite dosage form containing desmopressin acetate for buccal administration. Int. J. Pharm. X 2021, 3, 100082. [Google Scholar] [CrossRef] [PubMed]
  162. Zupančič, O.; Leonaviciute, G.; Lam, H.T.; Partenhauser, A.; Podričnik, S.; Bernkop-Schnürch, A. Development andin vitroevaluation of an oral SEDDS for desmopressin. Drug Deliv. 2016, 23, 2074–2083. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Enna, S.; Bylund, D.B. Desmopressin. In xPharm: The Comprehensive Phamcology Reference; Elsevier Inc.: Amsterdam, The Netherlands, 2007; pp. 1–2. [Google Scholar] [CrossRef]
  164. Suman, S.; Robinson, D.; Bhal, N.; Fraser, S.; MacCormick, A.; Williams, A.; Tadtayev, S. Management of nocturia: Overcoming the challenges of nocturnal polyuria. Br. J. Hosp. Med. 2019, 80, 517–524. [Google Scholar] [CrossRef] [PubMed]
  165. Aguirre, T.; Teijeiro-Osorio, D.; Rosa, M.; Coulter, I.; Alonso, M.; Brayden, D. Current status of selected oral peptide technologies in advanced preclinical development and in clinical trials. Adv. Drug Deliv. Rev. 2016, 106, 223–241. [Google Scholar] [CrossRef] [Green Version]
  166. Svensson, P.J.; Bergqvist, P.B.; Juul, K.V.; Berntorp, E. Desmopressin in treatment of haematological disorders and in prevention of surgical bleeding. Blood Rev. 2014, 28, 95–102. [Google Scholar] [CrossRef] [PubMed]
  167. Loomans, J.I.; Kruip, M.J.; Carcao, M.; Jackson, S.; van Velzen, A.S.; Peters, M.; Santagostino, E.; Platokouki, H.; Beckers, E.; Voorberg, J.; et al. Desmopressin in moderate hemophilia A patients: A treatment worth considering. Haematologica 2018, 103, 550–557. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Federici, A.B.; Mazurier, C.; Berntorp, E.; Lee, C.A.; Scharrer, I.; Goudemand, J.; Lethagen, S.; Nitu, I.; Ludwig, G.; Hilbert, L.; et al. Biologic response to desmopressin in patients with severe type 1 and type 2 von Willebrand disease: Results of a multicenter European study. Blood 2004, 103, 2032–2038. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Goldberg, N.; Nisenbaum, R.; Song, H.; Lillicrap, D.; Teitel, J.; James, P.; Sholzberg, M. Desmopressin responsiveness by age in type 1 von Willebrand disease. Res. Pract. Thromb. Haemost. 2020, 4, 1046–1052. [Google Scholar] [CrossRef] [PubMed]
  170. Pasalic, L.; Favaloro, E.J.; Curnow, J. Treatment of von Willebrand Disease. Semin. Thromb. Hemost. 2016, 42, 133–146. [Google Scholar] [CrossRef] [PubMed]
  171. Kaufmann, J.E.; Oksche, A.; Wollheim, C.B.; Günther, G.; Rosenthal, W.; Vischer, U.M. Vasopressin-induced von Willebrand factor secretion from endothelial cells involves V2 receptors and cAMP. J. Clin. Investig. 2000, 106, 107–116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Kaufmann, J.E.; Vischer, U.M. Cellular mechanisms of the hemostatic effects of desmopressin (DDAVP). J. Thromb. Haemost. 2003, 1, 682–689. [Google Scholar] [CrossRef] [PubMed]
  173. North, W.G.; Fay, M.J.; Longo, K.A.; Du, J. Expression of all known vasopressin receptor subtypes by small cell tumors implies a multifaceted role for this neuropeptide. Cancer Res. 1998, 58, 1866–1871. [Google Scholar] [PubMed]
  174. North, W.G. Gene regulation of vasopressin and vasopressin receptors in cancer. Exp. Physiol. 2000, 85, 27–40. [Google Scholar] [CrossRef]
  175. Petit, T.; Davidson, K.K.; Lawrence, R.A.; Von Hoff, D.D.; Izbicka, E. Neuropeptide receptor status in human tumor cell lines. Anti-Cancer Drugs 2001, 12, 133–136. [Google Scholar] [CrossRef] [PubMed]
  176. North, W.G.; Fay, M.J.; Du, J. MCF-7 breast cancer cells express normal forms of all vasopressin receptors plus an abnormal V2R. Peptides 1999, 20, 837–842. [Google Scholar] [CrossRef]
  177. Ripoll, G.V.; Garona, J.; Pifano, M.; Farina, H.G.; Gomez, D.E.; Alonso, D.F. Reduction of tumor angiogenesis induced by desmopressin in a breast cancer model. Breast Cancer Res. Treat. 2013, 142, 9–18. [Google Scholar] [CrossRef] [Green Version]
  178. Gately, S.; Twardowski, P.; Stack, M.S.; Cundiff, D.L.; Grella, D.; Castellino, F.J.; Enghild, J.J.; Kwaan, H.C.; Lee, F.; Kramer, R.A.; et al. 139 The mechanism of cancer-mediated conversion of plasminogen to the angiogenesis inhibitor angiostatin. Fibrinolysis Proteolysis 1997, 11, 39. [Google Scholar] [CrossRef]
  179. Alonso, D.F.; Skilton, G.; Farías, E.F.; Joffé, E.B.D.K.; Gomez, D.E. Antimetastatic effect of desmopressin in a mouse mammary tumor model. Breast Cancer Res. Treat. 1999, 57, 271–275. [Google Scholar] [CrossRef] [PubMed]
  180. Giron, S.; Tejera, A.M.; Ripoll, G.V.; Gomez, D.E. Desmopressin inhibits lung and lymph node metastasis in a mouse mammary carcinoma model of surgical manipulation. J. Surg. Oncol. 2002, 81, 38–44. [Google Scholar] [CrossRef] [PubMed]
  181. Hermo, G.A.; Torres, P.; Ripoll, G.V.; Scursoni, A.M.; Gomez, D.E.; Alonso, D.F.; Gobello, C. Perioperative desmopressin prolongs survival in surgically treated bitches with mammary gland tumours: A pilot study. Vet. J. 2007, 178, 103–108. [Google Scholar] [CrossRef] [PubMed]
  182. Hermo, G.A.; Turic, E.; Angelico, D.; Scursoni, A.M.; Gomez, D.E.; Gobello, C.; Alonso, D.F. Effect of Adjuvant Perioperative Desmopressin in Locally Advanced Canine Mammary Carcinoma and its Relation to Histologic Grade. J. Am. Anim. Hosp. Assoc. 2011, 47, 21–27. [Google Scholar] [CrossRef] [Green Version]
  183. Weinberg, R.S.; Grecco, M.O.; Ferro, G.S.; Seigelshifer, D.J.; Perroni, N.V.; Terrier, F.J.; Sánchez-Luceros, A.; Maronna, E.; Sánchez-Marull, R.; Frahm, I.; et al. A phase II dose-escalation trial of perioperative desmopressin (1-desamino-8-d-arginine vasopressin) in breast cancer patients. SpringerPlus 2015, 4, 428. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Ozgönenel, B.; Rajpurkar, M.; Lusher, J.M. How do you treat bleeding disorders with desmopressin? Postgrad. Med. J. 2007, 83, 159–163. [Google Scholar] [CrossRef]
  185. Terraube, V.; Marx, I.; Denis, C. Role of von Willebrand factor in tumor metastasis. Thromb. Res. 2007, 120, S64–S70. [Google Scholar] [CrossRef]
  186. Iseas, S.; Roca, E.L.; O’Connor, J.M.; Eleta, M.; Sanchez-Luceros, A.; Di Leo, D.; Tinelli, M.; Fara, M.L.; Spitzer, E.; Demarco, I.A.; et al. Administration of the vasopressin analog desmopressin for the management of bleeding in rectal cancer patients: Results of a phase I/II trial. Investig. New Drugs 2020, 38, 1580–1587. [Google Scholar] [CrossRef] [Green Version]
  187. Sobol, N.T.; Solernó, L.M.; Beltrán, B.; Vásquez, L.; Ripoll, G.V.; Garona, J.; Alonso, D.F. Anticancer activity of repurposed hemostatic agent desmopressin on AVPR2expressing human osteosarcoma. Exp. Ther. Med. 2021, 21, 566. [Google Scholar] [CrossRef]
  188. Hoffman, A.; Sasaki, H.; Roberto, D.; Mayer, M.J.; Klotz, L.; Venkateswaran, V. Effect of Combination therapy of Desmopressin and Docetaxel on prostate cancer cell DU145 proliferation, migration and growth. J. Cancer Biol. Ther. 2016, 2. [Google Scholar] [CrossRef] [Green Version]
  189. Bass, R.; Roberto, D.; Wang, D.Z.; Cantu, F.P.; Mohamadi, R.M.; Kelley, S.O.; Klotz, L.; Venkateswaran, V. Combining Desmopressin and Docetaxel for the Treatment of Castration-Resistant Prostate Cancer in an Orthotopic Model. Anticancer Res. 2019, 39, 113–118. [Google Scholar] [CrossRef] [PubMed]
  190. Vignon, P.; Laterre, P.-F.; Daix, T.; François, B. New Agents in Development for Sepsis: Any Reason for Hope? Drugs 2020, 80, 1751–1761. [Google Scholar] [CrossRef] [PubMed]
  191. Yao, R.-Q.; Xia, D.-M.; Wang, L.-X.; Wu, G.-S.; Zhu, Y.-B.; Zhao, H.-Q.; Liu, Q.; Xia, Z.-F.; Ren, C.; Yao, Y.-M. Clinical Efficiency of Vasopressin or Its Analogs in Comparison with Catecholamines Alone on Patients with Septic Shock: A Systematic Review and Meta-Analysis. Front. Pharmacol. 2020, 11, 563. [Google Scholar] [CrossRef] [PubMed]
  192. Selepressin Evaluation Programme for Sepsis-Induced Shock—Adaptive Clinical Trial—Full Text View—ClinicalTrials.gov. Available online: https://clinicaltrials.gov/ct2/show/NCT02508649 (accessed on 19 December 2021).
  193. Russell, J.A.; Vincent, J.-L.; Kjølbye, A.L.; Olsson, H.; Blemings, A.; Spapen, H.; Carl, P.; Laterre, P.-F.; Grundemar, L. Selepressin, a novel selective vasopressin V1A agonist, is an effective substitute for norepinephrine in a phase IIa randomized, placebo-controlled trial in septic shock patients. Crit. Care 2017, 21, 213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Ribot, S.; Abramowitz, S.; Kelhoffer, W.S.; Green, H.; Small, M.J.; Schwartz, I. CARDIOVASCULAR EFFECTS OF PHENYLALANYL-2-LYSYL-8-VASOPRESSIN (PLV-2). Am. J. Med Sci. 1963, 246, 479–484. [Google Scholar] [CrossRef] [PubMed]
  195. Cecanho, R.; de Luca, L.A.; Ranali, J. Cardiovascular Effects of Felypressin. Anesth. Prog. 2006, 53, 119–125. [Google Scholar] [CrossRef]
  196. Brahma, A.K.; Pemberton, C.J.; Ayeko, M.; Morgan, L.H. Single medial injection peribulbar anaesthesia using prilocaine. Anaesthesia 1994, 49, 1003–1005. [Google Scholar] [CrossRef] [PubMed]
  197. Yagiela, J.A. Vasoconstrictor agents for local anesthesia. Anesth. Prog. 1995, 42, 116–120. [Google Scholar]
  198. Meechan, J.G.; Cole, B.; Welbury, R.R. The influence of two different dental local anaesthetic solutions on the haemo-dynamic responses of children undergoing restorative dentistry: A randomised, single-blind, split-mouth study. Br. Dent. J. 2001, 190, 502–504. [Google Scholar] [CrossRef] [Green Version]
  199. Meechan, J.G. The effects of dental local anaesthetics on blood glucose concentration in healthy volunteers and in patients having third molar surgery. Br. Dent. J. 1991, 170, 373–376. [Google Scholar] [CrossRef]
  200. Volpato, M.C.; Ranali, J.; Amaral, I.M.; Demétrio, C.G.B.; Chalita, L.V. Acute toxicity (LD50 and CD50) of lidocaine and prilocaine in combination with adrenaline and felypressin. Indian J. Dent. Res. 1999, 10, 138–144. [Google Scholar]
  201. Sunada, K.; Nakamura, K.; Yamashiro, M.; Sumitomo, M.; Furuya, H. Clinically safe dosage of felypressin for patients with essential hypertension. Anesthesia Prog. 1996, 43, 108–115. [Google Scholar]
  202. Nery, D.T.F.; Ranali, J.; Barbosa, D.Z.; Júnior, H.M.; Pereira, R.M.A.; de Oliviera Afonso Pereira, P. Excisional Biopsy of the Pyogenic Granuloma in Very High-Risk Patient. Case Rep. Dent. 2018, 2018, 5180385. [Google Scholar] [CrossRef]
  203. Yamashita, K.; Kibe, T.; Shidou, R.; Kohjitani, A.; Nakamura, N.; Sugimura, M. Difference in the Effects of Lidocaine with Epinephrine and Prilocaine with Felypressin on the Autonomic Nervous System During Extraction of the Impacted Mandibular Third Molar: A Randomized Controlled Trial. J. Oral Maxillofac. Surg. 2020, 78, 215.e1–215.e8. [Google Scholar] [CrossRef]
  204. Kyosaka, Y.; Owatari, T.; Inokoshi, M.; Kubota, K.; Inoue, M.; Minakuchi, S. Cardiovascular Comparison of 2 Types of Local Anesthesia with Vasoconstrictor in Older Adults: A Crossover Study. Anesthesia Prog. 2019, 66, 133–140. [Google Scholar] [CrossRef]
  205. Ojeda-Yuren, A.S.; Cerda-Reyes, E.; Herrero-Maceda, M.R.; Castro-Narro, G.; Piano, S. An Integrated Review of the Hepatorenal Syndrome. Ann. Hepatol. 2021, 22, 100236. [Google Scholar] [CrossRef]
  206. Busk, T.M.; Bendtsen, F.; Møller, S. Hepatorenal syndrome in cirrhosis: Diagnostic, pathophysiological, and therapeutic aspects. Expert Rev. Gastroenterol. Hepatol. 2016, 10, 1153–1161. [Google Scholar] [CrossRef]
  207. Freeman, J.G.; Cobden, I.; Record, C.O. Placebo-Controlled Trial of Terlipressin (Glypressin) in the Management of Acute Variceal Bleeding. J. Clin. Gastroenterol. 1989, 11, 58–60. [Google Scholar] [CrossRef]
  208. Holmes, C.L.; Patel, B.M.; Russell, J.A.; Walley, K.R. Physiology of Vasopressin Relevant to Management of Septic Shock. Chest 2001, 120, 989–1002. [Google Scholar] [CrossRef] [Green Version]
  209. Krag, A.; Hobolth, L.; Møller, S.; Bendtsen, F. Hyponatraemia during terlipressin therapy. Gut 2010, 59, 417–418. [Google Scholar] [CrossRef] [Green Version]
  210. Solà, E.; Lens, S.; Guevara, M.; Martín-Llahí, M.; Fagundes, C.; Pereira, G.; Pavesi, M.; Fernández, J.; González-Abraldes, J.; Escorsell, A.; et al. Hyponatremia in patients treated with terlipressin for severe gastrointestinal bleeding due to portal hypertension. Hepatology 2010, 52, 1783–1790. [Google Scholar] [CrossRef]
  211. Krag, A.; Bendtsen, F.; Pedersen, E.B.; Holstein-Rathlou, N.-H.; Møller, S. Effects of terlipressin on the aquaretic system: Evidence of antidiuretic effects. Am. J. Physiol. Physiol. 2008, 295, F1295–F1300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Leithead, J.A.; Hayes, P.C.; Ferguson, J.W. Review article: Advances in the management of patients with cirrhosis and portal hypertension-related renal dysfunction. Aliment. Pharmacol. Ther. 2014, 39, 699–711. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Kusano, E.; Tian, S.; Umino, T.; Tetsuka, T.; Ando, Y.; Asano, Y. Arginine vasopressin inhibits interleukin-1β-stimulated nitric oxide and cyclic guanosine monophosphate production via the V1 receptor in cultured rat vascular smooth muscle cells. J. Hypertens. 1997, 15, 627–632. [Google Scholar] [CrossRef] [PubMed]
  214. Wakatsuki, T.; Nakaya, Y.; Inoue, I. Vasopressin modulates K(+)-channel activities of cultured smooth muscle cells from porcine coronary artery. Am. J. Physiol. Circ. Physiol. 1992, 263, H491–H496. [Google Scholar] [CrossRef]
  215. Hansen, E.F.; Strandberg, C.; Højgaard, L.; Madsen, J.; Henriksen, J.H.; Schroeder, T.; Becker, U.; Bendtsen, F. Splanchnic haemodynamics after intravenous terlipressin in anaesthetised healthy pigs. J. Hepatol. 1999, 30, 503–510. [Google Scholar] [CrossRef]
  216. Pliška, V.; Chard, T.; Rudinger, J.; Forsling, M.L. In vivo activation of synthetic hormonogens of lysine vasopressin: Nα-glycyl-glycyl-glycyl-[8-lysine]vasopressin in the cat. Eur. J. Endocrinol. 1976, 81, 474–481. [Google Scholar] [CrossRef] [PubMed]
  217. Döhler, K.D.; Meyer, M. Vasopressin analogues in the treatment of hepatorenal syndrome and gastrointestinal haemorrhage. Best Pract. Res. Clin. Anaesthesiol. 2008, 22, 335–350. [Google Scholar] [CrossRef] [PubMed]
  218. Nevens, F.; van Steenbergen, W.; Yap, S.H.; Fevery, J. Assessment of variceal pressureby continous non-invasive endoscopic registration: A placebo controlled evaluation of the effect of terlipressin and octerotide. Gut 1996, 38, 129–134. [Google Scholar] [CrossRef] [PubMed]
  219. Møller, S.; Hansen, E.F.; Becker, U.; Brinch, K.; Henriksen, J.H.; Bendtsen, F. Central and systemic haemodynamic effects of terlipressin in portal hypertensive patients. Liver Int. 2000, 20, 51–59. [Google Scholar] [CrossRef] [PubMed]
  220. Ioannou, G.N.; Doust, J.; Rockey, D.C. Terlipressin for acute esophageal variceal hemorrhage. Cochrane Database Syst. Rev. 2003, 2003, CD002147. [Google Scholar] [CrossRef] [PubMed]
  221. Dell’Era, A.; de Franchis, R.; Iannuzzi, F. Acute variceal bleeding: Pharmacological treatment and primary/secondary prophylaxis. Best Pract. Res. Clin. Gastroenterol. 2008, 22, 279–294. [Google Scholar] [CrossRef] [PubMed]
  222. Wróblewski, E.; Dabrowski, A. Management of variceal haemorrhage. Gastroenterol. Rev. 2010, 3, 123–144. [Google Scholar] [CrossRef]
  223. Sridharan, K.; Sivaramakrishnan, G. Vasoactive agents for the management of variceal bleeding: A mixed treatment comparison network meta-analysis and trial sequential analysis of randomized clinical trials. Drug Res. 2019, 69, 487–495. [Google Scholar] [CrossRef] [Green Version]
  224. Xu, X.; Lin, S.; Yang, Y.; Chen, Y.; Liu, B.; Li, B.; Wu, Y.; Meng, F.; Zhu, Q.; Li, Y.; et al. Development of hyponatremia after terlipressin in cirrhotic patients with acute gastrointestinal bleeding: A retrospective multicenter observational study. Expert Opin. Drug Saf. 2020, 19, 641–647. [Google Scholar] [CrossRef] [PubMed]
  225. Meng, Q.; Dang, X.; Li, L.; Liu, Z.; Li, L.; Wang, H. Severe hyponatraemia with neurological manifestations in patients treated with terlipressin: Two case reports. J. Clin. Pharm. Ther. 2019, 44, 981–984. [Google Scholar] [CrossRef] [PubMed]
  226. Pan, X.; Zhou, Z.; Jin, X.; Shi, D. Clinical characteristics and risk factors of severe hyponatremia in cirrhotic patients treated with terlipressin. J. Clin. Pharm. Ther. 2020, 45, 191–198. [Google Scholar] [CrossRef] [PubMed]
  227. Wang, X.; Tian, Y.; Lin, H.; Zang, L. Preparation and characterization of PEGylated terlipressin. J. Appl. Polym. Sci. 2010, 116, 3220–3224. [Google Scholar] [CrossRef]
  228. Gaudino, R.; Orlandi, V.; Cavarzere, P.; Chinello, M.; Antoniazzi, F.; Cesaro, S.; Piacentini, G. Case Report: SARS-CoV-2 Infection in a Child with Suprasellar Tumor and Hypothalamic-Pituitary Failure. Front. Endocrinol. 2021, 12, 596654. [Google Scholar] [CrossRef] [PubMed]
  229. Al-Kuraishy, H.M.; Al-Gareeb, A.I.; Qusti, S.; Alshammari, E.M.; Atanu, F.O.; Batiha, G.E.-S. Arginine vasopressin and pathophysiology of COVID-19: An innovative perspective. Biomed. Pharmacother. 2021, 143, 112193. [Google Scholar] [CrossRef]
  230. Sheikh, A.B.; Javed, N.; Sheikh, A.A.E.; Upadhyay, S.; Shekhar, R. Diabetes Insipidus and Concomitant Myocarditis: A Late Sequelae of COVID-19 Infection. J. Investig. Med. High Impact Case Rep. 2021, 9, 1–4. [Google Scholar] [CrossRef]
  231. Maffucci, I.; Contini, A. In Silico Drug Repurposing for SARS-CoV-2 Main Proteinase and Spike Proteins. J. Proteome Res. 2020, 19, 4637–4648. [Google Scholar] [CrossRef]
  232. Ahmad, J.; Ikram, S.; Ahmad, F.; Rehman, I.U.; Mushtaq, M. SARS-CoV-2 RNA Dependent RNA polymerase (RdRp)—A drug repurposing study. Heliyon 2020, 6, e04502. [Google Scholar] [CrossRef]
Figure 1. Timeline for development of AVP peptide analogues [28,29,30,31,32,33,34,35].
Figure 1. Timeline for development of AVP peptide analogues [28,29,30,31,32,33,34,35].
Ijms 23 03068 g001
Figure 2. Structures of (a) arginine vasopressin (AVP) and (b) oxytocin (OXT). Differences in sequence are shown in blue and red.
Figure 2. Structures of (a) arginine vasopressin (AVP) and (b) oxytocin (OXT). Differences in sequence are shown in blue and red.
Ijms 23 03068 g002
Figure 3. Schematic presentation of AVP receptors, their position and role in the body.
Figure 3. Schematic presentation of AVP receptors, their position and role in the body.
Ijms 23 03068 g003
Figure 4. Non-peptide agonists and antagonists of AVP receptor.
Figure 4. Non-peptide agonists and antagonists of AVP receptor.
Ijms 23 03068 g004
Figure 5. Natural peptide analogues of AVP, (a) LVP and (b) phenypressin. Amino acids that are not present in AVP are shown in red.
Figure 5. Natural peptide analogues of AVP, (a) LVP and (b) phenypressin. Amino acids that are not present in AVP are shown in red.
Ijms 23 03068 g005
Figure 6. Structure of peptide LVP analogues. The modifications are shown in blue and red. d = deamination of N-terminal Cys (Cys1); Dbt = 3,5-dibromo-l-tyrosine; Thi = thienilalanine; diHPhe = dihydrophenylalanine; Abu = 4-α-aminobutyric acid (AABA); Cha = 1-amino-cyclopentanecarboxylic acid (cyclohexylalanine); Eda = ethylendiamine.
Figure 6. Structure of peptide LVP analogues. The modifications are shown in blue and red. d = deamination of N-terminal Cys (Cys1); Dbt = 3,5-dibromo-l-tyrosine; Thi = thienilalanine; diHPhe = dihydrophenylalanine; Abu = 4-α-aminobutyric acid (AABA); Cha = 1-amino-cyclopentanecarboxylic acid (cyclohexylalanine); Eda = ethylendiamine.
Ijms 23 03068 g006
Figure 7. Structures of selected synthetic peptide analogues of AVP. Modifications are shown in different colors. d = deamination of N-terminal cysteine (Cys1); diPhe = 3,3-diphenyl-l-alanine; d-diPhe = 3,3-diphenyl-d-alanine; l-1-Nal = l-1-napthylalanine; l-2-Nal = l-2-napthylalanine; Aic = 2-aminoindane-2-carboxylic acid; Apc = 1-aminocyclopentane-1-carboxylic acid; Sar= l-sarcosine; NMeAla= N-methyl-l-alanine; HNle = l-homonorleucine; HyLeu = l-hydroxynorleucine.
Figure 7. Structures of selected synthetic peptide analogues of AVP. Modifications are shown in different colors. d = deamination of N-terminal cysteine (Cys1); diPhe = 3,3-diphenyl-l-alanine; d-diPhe = 3,3-diphenyl-d-alanine; l-1-Nal = l-1-napthylalanine; l-2-Nal = l-2-napthylalanine; Aic = 2-aminoindane-2-carboxylic acid; Apc = 1-aminocyclopentane-1-carboxylic acid; Sar= l-sarcosine; NMeAla= N-methyl-l-alanine; HNle = l-homonorleucine; HyLeu = l-hydroxynorleucine.
Ijms 23 03068 g007
Figure 8. Synthetic peptide analogues of AVP. Modifications are shown in different colors. hGln = homoglutamine, Orn(i-Pr) = N-δ-iso-propyl-l-ornithine.
Figure 8. Synthetic peptide analogues of AVP. Modifications are shown in different colors. hGln = homoglutamine, Orn(i-Pr) = N-δ-iso-propyl-l-ornithine.
Ijms 23 03068 g008
Figure 9. Analogues of dDAVP. Modifications are shown in different colors. Aic = 2-aminoindane-2-carboxylic acid; diPhe = 3,3-diphenyl-l-alanine; Thi = thienilalanine.
Figure 9. Analogues of dDAVP. Modifications are shown in different colors. Aic = 2-aminoindane-2-carboxylic acid; diPhe = 3,3-diphenyl-l-alanine; Thi = thienilalanine.
Ijms 23 03068 g009
Figure 10. AVP receptor antagonists.
Figure 10. AVP receptor antagonists.
Ijms 23 03068 g010
Table 4. Synthetic analogues of AVP. Amino acid residues that are not present in AVP are marked in bold.
Table 4. Synthetic analogues of AVP. Amino acid residues that are not present in AVP are marked in bold.
AnalogueSequenceFunctionsRefs.
Desmopressin, dDAVPdCys(&)-Tyr-Phe-Gln-Asn-Cys(&)-Pro-d-Arg-Gly-NH2 antidiuretic effect, increases plasma osmolality [64,120,121,122]
Selepressin Cys(&)-Phe-Ile-hGln-Asn-Cys(&)-Pro-Orn(iPr)-Gly-NH2 applied in septic shock [125,126,127]
FelypressinCys(&)-Phe-Phe-Gln-Asn-Cys(&)-Pro-Lys-Gly-NH2 vasoconstricting agent, used as an additive in anesthesia during dental procedures[128,129,130,131]
TerlipressinGly-Gly-Gly-Cys(&)-Tyr-Phe-Gln-Asn-Cys(&)-Pro-Lys-Gly-NH2treats bleeding caused by esophageal varices[132,133,134,135]
OrnipressinCys(&)-Tyr-Phe-Gln-Asn-Cys(&)-Pro-Orn-Gly-NH2vasoconstricting agent during myomectomy; in cirrhosis, as hepatorenal treatment[136,137,138]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Glavaš, M.; Gitlin-Domagalska, A.; Dębowski, D.; Ptaszyńska, N.; Łęgowska, A.; Rolka, K. Vasopressin and Its Analogues: From Natural Hormones to Multitasking Peptides. Int. J. Mol. Sci. 2022, 23, 3068. https://doi.org/10.3390/ijms23063068

AMA Style

Glavaš M, Gitlin-Domagalska A, Dębowski D, Ptaszyńska N, Łęgowska A, Rolka K. Vasopressin and Its Analogues: From Natural Hormones to Multitasking Peptides. International Journal of Molecular Sciences. 2022; 23(6):3068. https://doi.org/10.3390/ijms23063068

Chicago/Turabian Style

Glavaš, Mladena, Agata Gitlin-Domagalska, Dawid Dębowski, Natalia Ptaszyńska, Anna Łęgowska, and Krzysztof Rolka. 2022. "Vasopressin and Its Analogues: From Natural Hormones to Multitasking Peptides" International Journal of Molecular Sciences 23, no. 6: 3068. https://doi.org/10.3390/ijms23063068

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop