Next Article in Journal
Stimulation of Angiotensin II Receptor Subtype 2 Reduces Preeclampsia-like Symptoms in a Mouse Model of Preeclampsia
Previous Article in Journal
Unraveling the Mechanisms of S100A8/A9 in Myocardial Injury and Dysfunction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Agonists, Antagonists and Receptors of Somatostatin: Pathophysiological and Therapeutical Implications in Neoplasias

by
Argyrios Periferakis
1,2,3,
Georgios Tsigas
1,
Aristodemos-Theodoros Periferakis
1,2,
Carla Mihaela Tone
1,
Daria Alexandra Hemes
1,
Konstantinos Periferakis
3,4,
Lamprini Troumpata
1,
Ioana Anca Badarau
1,
Cristian Scheau
1,5,*,
Ana Caruntu
6,7,
Ilinca Savulescu-Fiedler
8,9,*,
Constantin Caruntu
1,10 and
Andreea-Elena Scheau
11
1
Department of Physiology, The “Carol Davila” University of Medicine and Pharmacy, 050474 Bucharest, Romania
2
Elkyda, Research & Education Centre of Charismatheia, 17675 Athens, Greece
3
Akadimia of Ancient Greek and Traditional Chinese Medicine, 16675 Athens, Greece
4
Pan-Hellenic Organization of Educational Programs, 17236 Athens, Greece
5
Department of Radiology and Medical Imaging, “Foisor” Clinical Hospital of Orthopaedics, Traumatology and Osteoarticular TB, 030167 Bucharest, Romania
6
Department of Oral and Maxillofacial Surgery, The “Carol Davila” Central Military Emergency Hospital, 010825 Bucharest, Romania
7
Department of Oral and Maxillofacial Surgery, Faculty of Dental Medicine, “Titu Maiorescu” University, 031593 Bucharest, Romania
8
Department of Internal Medicine, The “Carol Davila” University of Medicine and Pharmacy, 050474 Bucharest, Romania
9
Department of Internal Medicine and Cardiology, Coltea Clinical Hospital, 030167 Bucharest, Romania
10
Department of Dermatology, “Prof. N.C. Paulescu” National Institute of Diabetes, Nutrition and Metabolic Diseases, 011233 Bucharest, Romania
11
Department of Radiology and Medical Imaging, Fundeni Clinical Institute, 022328 Bucharest, Romania
*
Authors to whom correspondence should be addressed.
Curr. Issues Mol. Biol. 2024, 46(9), 9721-9759; https://doi.org/10.3390/cimb46090578
Submission received: 31 July 2024 / Revised: 29 August 2024 / Accepted: 31 August 2024 / Published: 2 September 2024

Abstract

:
Somatostatin is a peptide that plays a variety of roles such as neurotransmitter and endocrine regulator; its actions as a cell regulator in various tissues of the human body are represented mainly by inhibitory effects, and it shows potent activity despite its physiological low concentrations. Somatostatin binds to specific receptors, called somatostatin receptors (SSTRs), which have different tissue distributions and associated signaling pathways. The expression of SSTRs can be altered in various conditions, including tumors; therefore, they can be used as biomarkers for cancer cell susceptibility to certain pharmacological agents and can provide prognostic information regarding disease evolution. Moreover, based on the affinity of somatostatin analogs for the different types of SSTRs, the therapeutic range includes conditions such as tumors, acromegaly, post-prandial hypotension, hyperinsulinism, and many more. On the other hand, a number of somatostatin antagonists may prove useful in certain medical settings, based on their differential affinity for SSTRs. The aim of this review is to present in detail the principal characteristics of all five SSTRs and to provide an overview of the associated therapeutic potential in neoplasias.

1. Introduction

Somatostatin (SST), also called Somatotropin-Release-Inhibiting Factor (SRIF) or Growth Hormone Release Inhibitory Factor (GHRIF) was originally discovered in 1973 [1], and occurs in two forms, SRIF-14 and SRIF-28. Both molecules originate from the same initial precursor molecule, called preproSRIF, via post-translational processing [2,3,4]. Both isoforms are expressed in the relevant tissues, although at this point whether they are secreted by the same or different cells has not been elucidated [5].
After the initial discovery of somatostatin, and the elucidation of its inhibitory action on growth hormone (GH) secretion [1], it was rapidly realized that it was also associated with a number of other physiological functions [6,7]. It was found that somatostatin also inhibited thyrotropin (TSH) and was capable of suppressing its secretion in TSH-producing tumors; moreover, somatostatin was capable of inhibiting adrenocorticotropic hormone (ACTH) release in conditions associated with increased secretion [8]. However, only after documenting the existence of different somatostatin receptors (SSTRs) [9], it became possible not only to begin identifying all the different effects of somatostatin—a process continuing to this day—but also to begin correlating them with differential receptor expression in different cells, tissues and organs.
Through its receptors, somatostatin exhibits mostly inhibiting effects, via decreasing intracellular calcium and cyclic adenosine monophosphate (cAMP), while increasing potassium outflow [10]. For this reason, it is generally regarded as an inhibitory peptide [11]. The actions of somatostatin are endocrine, paracrine, autocrine and neuro-modulatory [12,13,14,15]; in these roles, somatostatin is expressed and active in the central nervous system and the gastrointestinal system exhibiting a variety of effects [16,17,18,19].
Despite all the different and multifaceted actions and effects of somatostatin, it is practically useless in therapy due to its extremely short half-life [20]. This led to the need to develop SRIF analogs, with longer half-lives and, preferably, higher affinities for certain SSTRs, compared to the physiological endogenous peptide [21]. The analogs in clinical use today are octreotide, lanreotide and pasireotide, while other compounds are also under consideration [22]. On the other hand, somatostatin antagonists have also been synthesized, and are still the subject of experiments and animal studies [23,24].
In this review, we aim to thoroughly document the functions and distribution of all five SSTRs and to highlight the clinical relevance of each receptor in neoplasms. Subsequently, we will present the common somatostatin analogs used in clinical practice, as well as the current research status and future perspectives of somatostatin antagonists.

2. Description of Somatostatin Receptors

2.1. Overview of Somatostatin Receptors

The discovery of somatostatin is attributed to Brazeau and Guillemin [1] who isolated the then-unknown peptide from the ovine hypothalamus. It was a considerable time after that the five distinct SSTRs were identified, owing to the fact that radioligand studies allowed only for the distinction between SRIF binding sites and did not reveal the existence of five distinct receptors. The receptors are designated as SST1, SST2, SST3, SST4, and SST5, respectively [5,25].
According to the results of the SRIF-affinity studies, the five receptors can be classified into two groups, based mostly on their pharmacological and structural profiles. These are the SRIF-1 group (including SSTR2, SSTR3, and SSTR5) and the SRIF-2 group which includes SSTR1 and SSTR4 [21,26]. Despite their similarities, each receptor has different molecular weights and is coded for by different genes, with particular tissue distribution (Table 1), and functions (Table 2).

2.2. Structure and Properties of Somatostatin Receptors

All five of the SSTRs are G protein-coupled receptors (GPCRs), belonging to the rhodopsin family; in general, GPCRs act as receptors for extracellular stimuli, both physical and chemical, and their notable pathophysiological correlations make them prime targets for pharmacological interventions [61]. In accordance with the characteristic architecture of the GPCRs, all SSTRs share the common structure of seven transmembrane segments [62], and a sequence similarity of up to 57% [5]. However, their amino acid sequences differ, particularly at the level of the third intracellular loop as well as at the C-terminal end, leading to specific functional features [63].
Apart from the case of SSTR2, all SSTRs are coded from intronless genes. Regarding SSTR2, there exist two subtypes, SSTR2A and SSTR2B, which are the product of alternative splicing; while both subtypes exist in rodents, only SSTR2A is found in humans [64]. On the other hand, there are two SSTR5 variants in humans, produced by a different process [65]. Another particularity of SSTR genes is that they do not have TATA boxes [66], which are conserved sequences in the gene promoter region [67]. This may indicate that the genes themselves are insensitive to the knockout of their promoter-binding transcription factors [68].
A further common aspect of SSTRs is their association with Gi/Go proteins, which are part of the pertussis-toxin (PTX) sensitive protein family [37,69]. SSTR activation invariably leads to the reduction in intracellular cAMP and Ca2+, through inhibition of adenylyl cyclase. They can also cause inhibition of inward rectifier current K+ channels (Kir3.x) leading to hyperpolarisation and inhibition of intracellular Ca2+ entry through voltage-gated Ca2+ channels [70]; such actions inhibit hormone secretion [5]. The study of Pan et al. [71], also mentioned the role of protein tyrosine phosphatases in inhibiting cell proliferation, although the complete spectrum of molecular signaling associated with SSTR-induced proliferation inhibition is still poorly understood [5]. SSTRs can also form homodimers and heterodimers which can alter their pharmacological properties and signaling pathways [72]. They can also heterodimerize with other GPCRs leading to complex interactions and additional functional roles for the particular cells [73].
Receptor desensitization can occur for SSTRs similar to other GPCRs via phosphorylation and uncoupling from the proteins [74]. Subsequently, receptor internalization takes place, which is essential for the resensitization process [75]. The internalization process occurs at different concentrations of SST or SSTR agonist and in different conditions for each cell type and tissue [15,72,76,77].

2.3. Localization, Functions and Signaling Pathways of Somatostatin Receptors

2.3.1. SSTR1

SSTR1, a highly glycosylated protein [31], is most prominently expressed in cells of the jejunum and the stomach [27]. It is also found in the CNS, concentrated in the presynaptic part of the neurons [28], where its activation leads to suppression of somatostatin secretion [78], of release-inhibiting hormones in general, and of growth-hormone-releasing hormone (GHRH) [79] and growth hormone (GH) itself [80] in particular. SSTR1 also acts as an inhibitory autoreceptor in the mediobasal hypothalamus, basal ganglia and retina [81]. Another role of SSTR1 is the regulation of the excitability of the cholinergic neurons of the forebrain [82]; an anti-inflammatory effect and a role in nociception have also been proposed [45,83,84]. The expression of SSTR1 in the β cells of Langerhans islets [85], most probably explains its role in regulating insulin secretion [86,87]. It is also found in the cell lining of veins and arteries [5].
Its signaling pathway, demonstrated to be highly influenced by the intracellular milieu, was found to be associated with G proteins both sensitive and insensitive to PTX [88,89,90,91,92]. Depending on the particular cell type, SSTR1 may exert its function via different signaling pathways [87,93,94]. SSTR1 activation has also been demonstrated to have cell migration-inhibiting properties, via inhibition of focal adhesion and actin stress fiber formation [95]. Activation of SSTR1 has also been shown to inhibit cell growth by tyrosine kinase dephosphorylation [96] and by p21 upregulation [97].

2.3.2. SSTR2

SSTR2 is also highly glycosylated, and although two subtypes exist in nature, SSTR2A and SSTR2B, only the first is expressed in humans [64]. It is commonly found in central and myenteric neurons, neuroendocrine cells of the gastric antrum, anterior pituitary cells and pancreatic islets [33,98,99,100] and also cells at the deeper layers of the cerebral cortex and in the cerebellum [101,102,103]. In general, SSTR2 is the most commonly detected SSTR [5], being also located in peripheral organs and structures [33,43,104,105]. Finally, SSTR2 is involved in a rather peculiar intracellular recycling pathway, in some cell lines, but the implications of this phenomenon are not fully understood [5]. This receptor is instrumental in the secretion of GH and thyroid-stimulating hormone (TSH) [50], is involved in inhibitory neuromodulation [5] and has a potential anti-epileptic action [102]. Feeding and drinking behaviors are also associated with SSTR2 [51,106] and the reduction in stress-induced or stress-related endocrine functions of the CNS and the pituitary gland [52,60]; an anti-depression role has also been proposed, associated with activation of SSTR2 and SSTR3, which in turn modulates serotonin release, although the precise mechanisms are still uncertain [47,48].
The signaling mechanisms are similar to that of SSTR1 and likewise, the effects are similar [5]. However, in some cell types, the inhibition of cAMP production by SSTR2 proved impossible [107]; this may be important in elucidating structural and functional differences in the morphofunctional differences of the G-protein-coupling domains of SSTRs. SSTR2 is involved in the regulation of the mitogen-activated kinase (MAP) kinase pathway in pituitary cells [108]; the tyrosine kinase pathway is also activated by SSTR2 [109,110]. The effects of SSTR2 activation are associated with cell cycle arrest and, by implication, inhibition of multiplication [111,112]. There is also a noteworthy interaction between SSTR2, in some cells, and Shank proteins, as mentioned by Günther et al. [5]; these proteins may be implicated in the regulation of synaptic transmission [113].
In some types of tumor cells, the activation of SSTR2 can trigger apoptosis [50]. It has also been reported that the dephosphorylation of SSTR2 is associated with reduced cell migration and invasion [114]. Notably, there is a difference in the genetic and epigenetic makeup of SSTR2 genes in some tumor cell lines [115,116]. The phosphorylation of specific residues of threonine and serine of this receptor at the carboxyl-terminal tail is related to its internalization in some cell types [117].

2.3.3. SSTR3

SSTR3 is similar to SSTR2, being over 40% homologous; it is again heavily glycosylated in vivo [36]; despite its relative similarity to SSTR2, it is characterized by some unique structural features, compared to other SSTRs, such as the presence of a significantly longer carboxyl-terminal tail which lacks a potential palmitoylation locus [5]. The activation of SSTR3 results in adenylyl cyclase inhibition [118], potassium channel currents [119] and calcium channel currents activation [54]. In CHO-K1 cells, its activation also resulted in an increase in p53 and Bax levels [120], a fact that may prove useful in inducing apoptosis or preventing the growth of tumor cells. SSTR3 is also rapidly downregulated after prolonged agonist exposure, compared with other SSTRs [118,121]. The prime function of SSTR3 at a cellular level is hormone release inhibition; common examples include GH secretion inhibition [55] and insulin secretion inhibition [54]. Examples in rodents suggest that the presence of SSTR3 in brain nerve cells might be associated with memory and object recognition [122] and an anticonvulsant effect [123].

2.3.4. SSTR4

This receptor is predominantly expressed in the brain and is found in a number of areas, such as the hippocampus, the hilar region of the dentate gyrus, amygdala, and hypothalamus [5]; its presence has also been ascertained in retinal ganglion cells [124]. It was also detected in pulmonary, heart and placental cells [125,126]. A host of other organs and structures are also characterized by SSTR4 expression [40]. Available data suggest that this receptor is not important in controlling the function of the anterior pituitary [127,128], but, based on mice studies, it is important in memory-related processes [58], locomotor functions [57,129], and stress-responses [130,131]. The activation of SSTR4, which in turn inhibits Kir3.x and voltage-activated Ca2+ channels is presumably associated with analgesic effects [132].

2.3.5. SSTR5

The sequence of this receptor is less conserved among species compared to the other four SSTRs [5]. The signaling pathways of SSTR5 are common with those mentioned for other SSTRs, such as inhibition of cAMP [15,133,134,135,136]. However, this receptor is associated with a number of different signaling mechanisms which are dependent on the biochemical conditions in the environment of the receptor [15,136]. SSTR5 activation can lead to intracellular Ca2+ increase by activation of phospholipase C [137,138] but also to intracellular Ca2+ decrease by inhibition of voltage-dependent Ca2+ channels [139]. A number of specific pathways have been described by a number of researchers, including a cyclic guanosine monophosphate-dependent pathway, Gq-mediated mitogen-activated protein kinase activation, and Gα-mediated stress-activated protein kinases (SAPK)/Jun amino-terminal kinases (JNK) pathway activation [134,135,140,141,142,143]. In general, SSTR5 has an inhibitory activity on secretion [50,144,145]. SSTR5 also exists in Sertoli cells [146], and in aortic smooth muscle cells [147] and has a limited CNS distribution [5].

2.4. Importance of Somatostatin Receptors in Tumoral Pathology

The use of SSTRs as targets for pharmacological intervention has been considered extensively in the last decades. As shall be presented, especially so in the case of neoplasias, their expression patterns are important for prediction and treatment purposes. In general, it can be said that tumors having a low expression of SSTRs are less susceptible to therapy. Most tumors express more than one SSTR; however, there is usually a predominance of one subtype of receptor over the others and this might constitute a useful feature in both diagnosis and treatment.

2.4.1. Somatostatin Receptors and Neuroendocrine Tumors

Neuroendocrine tumors (NETs) of the thorax are characterized by an increased SSTR2 expression [148], and sometimes of SSTR5 [149]. The expression of this receptor is also critical both for the diagnosis and treatment of small intestinal NETs and correlates positively with patient survival [150]. For NETs in general, Chang et al. [151] have argued in favor of a short-acting somatostatin diagnostic test. SSTR2 seems to be predominantly expressed also in the case of malignant insulinoma, although this did not predispose to a successful octreotide therapy [152]. Interestingly, in octreotide-treated patients, SSTR2 is internalized in the cell membranes of NETs [153]. Partial receptor internalization was noted in the case of pasireotide-treated patients [117,118].
The predominance of SSTR2 in the case of NETs has also been confirmed by other researchers [154,155,156]. In rectal NETs, an increased SSTR2 expression correlates with a better response to therapy and more favorable outcomes [157]. In more than half of NETs, SSTR3 expression can also be detected [158,159]; tumors from various regions, exhibiting neuroendocrine differentiation are not apparently characterized by elevated SSTR3 expression [160]. In the same manner, SSTR2 was the most important SSTR expressed in meningiomas [161,162], a fact that leaves open the exploration of the therapeutic potential of somatostatin analogs, at least in the case of inoperable metastatic tumors [163].
SSTR2 and SSTR3 were found to be expressed in pheochromocytomas and paragangliomas by Leijon et al. [164], with the other three SSTR receptors not being expressed at all; immunohistochemistry of SSTR2A should be taken into consideration for personalized treatment schemes in paragangliomas [165]. SSTR3 was found to be the most expressed receptor in a study also of Parvizi et al. [166]; curiously, all SSTRs were found to be expressed at different rates in paraganglioma cells in the study of Kaemmerer et al. [167]. Nevertheless, the expression of SSTR2 renders paragangliomas, along with a host of other tumors, responsive to octreotide radioligand therapy [168].
In the case of acromegaly, SSTR2 is an important receptor due to its expression on adenoma cell membranes [169,170]. Therefore, the high expression of SSTR2 may contribute to the response to therapy [171,172,173]. Histologically, it is possible to divide adenomas into two different types, densely and sparsely granulated ones [174,175]; the first type, expresses SSTR2 much more than the latter and is thus more responsive to treatments with SSTR ligands [176,177]. It must be noted that, though not the most important SSTR in that case, SSTR5 appears to play a modulatory role; when expressed, it seems to upregulate the SSTR2-associated pathways, thus improving the therapeutical effectiveness, when ligands of SSTRs are used [44].
SSTR2 is also the one most predominantly expressed in thymic tumors [178]. The importance of SSTRs in thymic tumors is still a matter of debate, as there is a difference between in vitro cultured cells and in vivo thymomas; SSTR3 seems to be the only one expressed in both cases, while SSTR1 and SSTR2 were detected only in vivo [179]. In general, for thymomas, SSTR3 is the predominant receptor, and it is presumed that the absence of endogenous somatostatin, and hence its permanent inactive state, leads to uncontrolled proliferation [179].
In the cases of insulinomas [180], SSTR2 is the most commonly expressed receptor, although SSTR3 generally correlates with a larger tumor size and SSTR3 and SSTR5 co-expression correlates with a less positive outcome; in general, SSTR expression in insulinomas is weak, with about 20% of such tumors not expressing SSTRs at all [181].
In pituitary adenomas, which secrete GH, this receptor was found to be prominently expressed, along with SSTR5 and SSTR2 [36,182]; interestingly, in the case of pituitary adenomas secreting ACTH, SSTR2 expression was not usually detected [36,104]. In other pituitary tumors, only SSTR3 expression was notable [36,183,184].
In general, the available research data suggest that SSTR4 is not a common SSTR in neoplasias, although conflicting data regarding its expression on insulinomas exist [185,186]. The importance of SSTR2, and its common association with SSTR5 expression, was also highlighted in Merkel carcinoma cells [187,188,189]. The high SSTR2 expression has also been confirmed by Fagerstedt et al. [190], who have also noted its potential as a prognostic marker.

2.4.2. Somatostatin Receptors and Squamous Cell Carcinomas

SSTR5 was found to be the predominant receptor in head and neck squamous carcinoma cells [191], whereas SSTR1 and SSTR2 were expressed at high rates and SSTR3 and SSTR4 were seldomly expressed [191]. Conversely, Lum et al. identified concomitantly high expression of SSTR1 and SSTR2 while SSTR5 expression was found to be particularly low [192]. In general, head and neck squamous carcinomas are characterized by a complex pathophysiology [193] and SSTR expression patterns and implications in the signaling pathways and carcinogenesis are not clear [194,195,196].
On a different note, the methylation of the SSTR2 genes was found to be a poor prognostic factor in the case of laryngeal squamous cell carcinoma [116]; the expression of SSTR2 in the case of nasopharyngeal carcinoma was found to be moderate to high [197]. The importance of this receptor and its associated signaling, expression and related epigenetics in squamous cell carcinomas of the head and neck has represented a research interest for Fan et al. [198]. SSTR2 and SSTR5 expression were found to be high in the case of uveal melanoma cells [199]. Based on the study of Valsecchi et al. [200], it is possible to use SSTR expression both therapeutically and to estimate survival rates.

2.4.3. Somatostatin Receptors and Reproductive System Carcinomas

In breast cancer, SSTR1, SSTR2 and SSTR3 were highly expressed, with SSTR4 and SSTR5 at comparatively lower percentages, but still over 60% [201]. Another study [202], found SSTR1 expression to be pronounced, followed by SSTR4 and the rest at considerably lower levels. In the case of ovarian cancer, one study [203] determined that the expression of SSTR1, SSTR2 and SSTR5 was predominant, while another [204] found SSTR3 expression to be higher. In general, somatostatin expression and its clinical significance in ovarian tumors is a matter requiring further research [205]. A recent study, by Zhao et al. [206], found SSTR5 to be the receptor most expressed in ectopic endometrial tissue, followed by SSTR2 [207]. The treatments currently available are limited and do not target SSTRs [208].
For prostate cancer, SSTR1 was found to be the primary receptor expressed in the case of prostate neoplasias, amongst 80 different samples [209]; this overexpression of SSTR1 has been also documented by other researchers [210,211]. There it inhibits the production of prostate-specific antigen (PSA) and has antiproliferative effects [212]. Another study correlated the increased expression of SSTR1 with increased tumor suppression [213]. However, overall, SSTR expression is relatively low in prostate neoplasia, with SSTR5 and SSTR2 being expressed at a low percentage of about 10% [214,215].

2.4.4. Somatostatin Receptors and Digestive Carcinomas

In gastric cancer, SSTR2 is the most predominant [216], while SSTR3 seems to be more expressed, on average, in healthy and not in cancerous gastric mucosa [217]. It was also found that the expression of SSTR genes was inversely proportional to metastatic potential [218]. The expression of SSTR2 in colorectal cancer cells can be used as a prognostic marker, according to the study of Casini Raggi et al. [219], although it was noted that SSTR2 expression between healthy and cancer cells was not that much different.
SSTR5 was found to be expressed more than all other SSTRs in MALT-type lymphomas, and to be more expressed in gastric-type lymphomas, along with SSTR3 and SSTR4, compared to extragastric ones; in addition, the presence of SSTR5 in lymphomas was found to correlate with a more favorable prognosis [43].

3. Therapeutical Uses of Somatostatin Analogs in Neoplasias

In general, the activation of SSTRs, in the case of neoplasias, exerts effects at a cellular level—antiproliferative [220,221,222], apoptotic [223,224,225]—and at a systemic level—antineoplastic [45,226,227] and anti-inflammatory [22,228,229,230]. Therefore, the use of somatostatin analogs has seen widespread use, in different neoplastic syndromes; on the other hand, taking advantage of the inhibitory effects of somatostatin on hormone and bioactive molecules secretion, some use of somatostatin analogs has demonstrated promising effects in other pathological conditions.
The primary reason for using somatostatin analogs instead of somatostatin itself is that endogenous somatostatin is very rapidly degraded by the human body [5]; the degradation of somatostatin is performed by ubiquitous plasma and tissue peptidases [231]. Another problem with somatostatin use was revealed when attempting to manage acromegaly, the initial disease where somatostatin and its analogs saw therapeutic application; the suppression of GH secretion was accompanied by an unacceptable suppression of insulin secretion [232,233,234,235]. At first, it was difficult to explain how and why somatostatin analogs inhibited GH secretion more than insulin secretion, a conundrum only solved after the existence of different SSTRs was realized [9]. Current research interests include the identification of SSTR pathways, and their interference with physiologic regulation mechanisms [5]. This is an important parameter especially in the design of analogs for therapeutic purposes.
At some level, all somatostatin analogs mimic the actions of somatostatin, and this is their common pharmacodynamical property. However, they preferentially act on different receptors (Table 3), and therefore it is possible to choose different analogs to take advantage of the differential expression of SSTRs in different tumors. For example, while endogenous somatostatin has its greatest selectivity for SSTR3, with a Ki of about 6 nM, octreotide has a Ki of over 1000 nM for SSTR1 and SSTR5, lanreotide has a commensurate selectivity only for SSTR1, and pasireotide has a Ki of >100 nM for SSTR4 [9,92,236,237,238].
Of the existing somatostatin analogs, only three have been approved for clinical use, octreotide, lanreotide and pasireotide (Table 4); the two former are first-generation analogs, and the latter is a second-generation analog [239]. Octreotide, a cyclic octapeptide, is more stable from a metabolic point of view, compared to somatostatin, because of its D-confirmation of amino acids and the presence of a disulfide bridge [240]—these increase its resistance to degradation by intestinal peptidases [241] Nevertheless, it retains the basic Phe-Trp-Lys-Thr motif of somatostatin [242]. The secondary structure of lanreotide is essentially the same, but there are some differences in amino acids compared to octreotide [243]. Pasireotide is completely different in terms of composition and structure from the other two analogs, being a six-membered homodetic cyclic peptide composed from L-phenylglycyl, D-tryptophyl, L-lysyl, O-benzyl-L-tyrosyl, L-phenylalanyl and modified L-hydroxyproline residues joined in sequence [244]
Regarding the rest, vapreotide has had some applications in humans, animals and in vitro [245,246,247,248,249,250,251,252,253,254,255,256,257], but it has had limited application in neoplasias as of yet [258,259,260]. Veldoreotide is a recently synthesized analog and little research exists on its potential applications [261]. Finally, seglitide, existing since the 1980s [262], has not been widely used, with a few exceptions [9,246].
Table 3. Pharmacological properties of clinically used somatostatin analogs [263,264,265,266,267,268].
Table 3. Pharmacological properties of clinically used somatostatin analogs [263,264,265,266,267,268].
Somatostatin AnalogOctreotideLanreotidePasireotide
Most effective administration routeSubcutaneousSubcutaneousSubcutaneous
Cmax attainment30 min (subcutaneous)
1.67–2.5 h (oral)
n/a0.25–0.5 h
Volume of distribution (VL)13.6–30.4 L15.1 L>100 L
Protein binding65%n/a88%
MetabolismLiverGastrointestinal tractLiver
Half-life2.3–2.7 h22 d12 h
Clearance7–10 L/h23.1 L/h7.6 L/h
EliminationUrinary (32%)
Faecal (30–40%)
Urinary (<5%)
Faecal (<0.5%)
Hepatic (48.3%)
Urinary (7.63%)
Table 4. Use of somatostatin analogs in the treatment and management of various conditions.
Table 4. Use of somatostatin analogs in the treatment and management of various conditions.
Somatostatin AnalogConditionTumor-Expressed SSTRsReferences
OctreotideAcromegalySSTR2, SSTR5 (minor role)[269,270]
Congenital hyperinsulinismSSTR2[271]
Thymic tumorsSSTR2 (minor role), SSTR3[178]
Lymphoma (potential treatment)SSTR2[272]
Neuroendrocrine tumorsSSTR2, SSTR5 (minor role)[273]
Merkel cell carcinomaSSTR2, SSTR5[274,275]
Hepatocellular carcinomaAll SSTRs[21]
LanreotideAcromegalySSTR2[34]
Neuroendocrine tumorsSSTR2, SSTR5 (minor role)[273,276,277]
Hepatocellular carcinomaAll SSTRs[278]
PasireotideAcromegalySSTR2 (minor role), SSTR5[34,279]
Congenital hyperinsulinism (proposed) and drug-induce hyperglycaemia in Cushings syndromeSSTR5[271,280]
Neuroendrocrine tumorsSSTR1, SSTR2, SSTR3, SSTR5[281]
Despite the numerous potential uses of somatostatin analogs outlined above, their use is not without a risk of side-effects ranging from mild to severe and even life-threatening. Frequent side-effects may be localised (local infusion reactions, pain and swelling), nervous system-associated (headache, nausea, or vomiting), gastrointestinal (abdominal pain/bloating, diarrhea, and gallstone formation—the frequency of that depends on the presence of other comorbidities), and appetite loss. Other effects such as arrhythmias and/or bradycardia, hypertension, vitamin B12 deficiency, impaired glucose tolerance (depending on comorbidities) and fatigue/malaise are rare; finally, the very rare side effects comprise hypothyroidism, hepatic injury and pancreatitis [266,267,268,282].
For example, as analyzed in the following parts, in conjunction with octreotide, octreotide-induced hypoglycaemia, in patients with insulinomas, may even prove fatal [283]; likewise, the gallstone incidence increase may be so high, that even precautionary surgical removal of the gallbladder may be recommended [284].

3.1. Therapeutical Uses of Octreotide in Neoplasias

Octreotide was first synthesized in 1979 by W. Bauer [285], and compared to somatostatin, it is an even more potent inhibitor of glucagon, insulin and growth hormone release [14]. While initially reported as carrying a potential analgesic effect in cancer patients, this was disproved in a clinical trial [286]; another study reported that it may be beneficial in cases of chronic non-malignant pancreatic pain [287]. Today, octreotide is the oldest somatostatin analog in clinical practice, being used for over 40 years [288].
In general, octreotide inhibits growth hormone and glucagon, various gastrointestinal hormones and reduces splanchnic blood flow through mechanisms similar to somatostatin, but with higher effectiveness [264]. Common side effects comprise the reduction in gall bladder contractility, thyroid stimulating hormone release [264] and B12 vitamin levels [266]. Notably, octreotide can be found in breast milk, in high concentrations [289,290], but this is not considered very dangerous for the infant given the low oral absorption of the drug [263]. Extreme caution is required if and when octreotide is administered in elderly people, as it decreases hepatic and renal and even cardiac functions [291]. To our knowledge, there are no reports in scientific literature regarding octreotide overdose. On the other hand, octreotide administration is effective in treating sulfonylurea overdose [292,293,294,295]; this drug is the oldest antidiabetic medication, dating back to the 1950s [296], and is still used by type 2 diabetes patients [297,298,299,300].
A case report, regarding insulin glargine overdose in a type 2 diabetes patient [301], corroborates the potential of octreotide to prove useful in treating antidiabetic drug overdose. Finally, it has been theorized that octreotide could be administered to treat drug overdose in general [302]. However, alongside the therapeutical merits of octreotide, the cessation of its administration, especially if it is prolonged, may be associated with significant side effects [303,304,305].
Subcutaneous administration is the most effective, with virtually complete absorption, [263] while oral administration was found to be less efficient by approximately 30% [266]. Oral octreotide administration must be performed on an empty stomach, as the presence of food reduces the absorption of octreotide by about 90%. This is important in the context of bioavailability. For example, the action of octreotide on tumor cells typically extends for between 8 and 12 h post-administration [291].

3.1.1. Octreotide in the Treatment of Thymic Tumors

Thymic tumors are very rare, and they comprise a group of different histological patterns, which are in turn associated with different survival rates [178]. In general, in thymic cancers, more than one SSTR is present, with the expression of SSTR2 being predominant in the majority of cases [178]. Octreotide has been shown to regulate the development and maturation of thymocytes [306,307] and it is worth mentioning that the activation of SSTR1 and SSTR2 promotes the proliferation and maturation of thymocytes in fetuses [308,309]—this is one of the very few non-inhibitory actions of somatostatin in the human body. Completely different from the pattern of healthy cells, thymomas were found to mostly express SSTR3. In general, the use of octreotide to combat thymic tumors can be regarded as promising, based on a number of case studies [310,311,312,313,314,315,316] and case series [317,318,319]. However, reported administered doses vary significantly from 20 mg per month for long-acting release forms up to 200 mg three times per day [314,316], With various dosages available and the ensuing difficulty in assessing the side effects it is recommended for octreotide to be used only in cases where significant octreotide uptake by the tumor cells can be demonstrated [178,314].

3.1.2. Octreotide in the Management and Treatment of Neuroendocrine Tumors (NETs)

Neuroendocrine tumors (NETs) are a group of heterogeneous neoplasms associated with cells of the diffuse neuroendocrine system (DNES); different NET classification schemes have been proposed [320,321,322]. Regardless, their clinical significance very much depends on their site of origin [273]. Crucially, NETs are known to have a pronounced expression of SSTRs [282]; in most cases, SSTR2 is the most prominently expressed, at least in gastroenteropancreatic NETs [323]. In general, long-acting somatostatin analogs present many advantages in treating well-differentiated NETs [324]. Personalized NET treatment with somatostatin analogs presents one of the challenges of modern personalized medicine [325]. Identification of adequate biomarkers is necessary for refining personalized treatment in NETs, and microRNA may be useful response predictors and prognostic markers [325]. Reported dosages of long-acting analogs vary between 20 and 180 mg/28 days [324]. Recent research indicated that there are no substantial gender differences in response to somatostatin analogs treatment in NETs, despite early evidence to the contrary from preclinical studies [326].
Octreotide with its high affinity for SSTR2 can be effectively used to manage the symptoms associated with NETs, as evidenced by some relevant studies [284,327,328]. NETs are associated with the so-called carcinoid syndrome, frequently caused by the release of serotonin, and bioactive peptides by these tumors [329]; common symptoms comprise flushing and diarrhea, but atypical symptoms, such as pellagra and bronchospasms, have also been recorded [330]. In such cases, octreotide may be used to manage carcinoid crises or as prophylaxis [331], although there is no consensus as to its effectiveness [332,333]. Also, the heterogeneity as well as the absence of a sufficient amount of clinical data present therapeutic challenges in NETs [334].
A testament to the importance of SSTR2 in the case of NET treatment is the case of advanced insulinomas, which typically have a low SSTR2 expression and are therefore less or non-responsive to octreotide treatment, demonstrating that SSTR expression is a critical factor in treatment effectiveness [273]. In addition, octreotide treatment can lead to hypoglycemia, possibly because of the independent downregulation of glucagon [284]. It is also possible to use octreotide to reduce gastrin secretion in the case of Zollinger-Ellison syndrome [335]; this may be localized or associated with multiple endocrine neoplasia type 1 [336]—presumably, in this latter case octreotide or some other somatostatin analog could be used target multiple neoplasias.
In the case of neuroendocrine carcinoma (NEC) of the lung, as with various neuroendocrine tumors, they frequently express more SSTRs compared to normal tissues [337], and are thus susceptible to somatostatin action [338]. This strategy is frequently sufficient for symptom relief—symptoms arising from the overproduction of hormones from the tumors—but can rarely reduce tumor size [339]. But for any targeted somatostatin therapy, the precise pattern of SSTR expression of the tumor must be known [340]. This analysis may be performed by a variety of methods, mainly immunohistochemical [341,342,343,344,345,346,347,348], and is essential, in the case of all therapeutical somatostatin analog use and not just for octreotide.
For lung NECs, the predominant SSTRs are SSTR2A and SSTR1, with SSTR4 and SSTR5 less frequently expressed [349], depending on the degree of differentiation of the tumor [337,346]. In the case of small cell lung cancer, a type of high-grade neuroendocrine cancer [350], the presence of SSTRs may render them responsive to somatostatin analog treatment, although it has, as of yet, no known predictive value [351]. The determination of SSTR expression may be possibly performed in the future with next-generation sequencing as proposed by Kruglyak et al. [352] and Cainap et al. [353]; in fact, such sequencing could be applied to most tumors [354], specifically to look for SSTR expression or mutations.
Based on current evidence, it is possible that in slow-growing diseases, treatment with somatostatin analogs is not really more effective than the “watch-and-wait” approach [273]. This was noted in 1996, by Perry and Vinik [355], who commented that for slow-progressing tumors, it was difficult to assess the effectiveness of octreotide therapy on tumor growth; nevertheless, there was a clear therapeutical effect, in cases of metastatic tumors. On the other hand, again taking advantage of the increased expression of SSTRs observed in NETs, radiolabelled octreotide can be used for systemic radiotherapy, to deliver radioisotopes specifically to tumor cells [356,357,358,359,360]. At any rate, the administration of octreotide, or lanreotide, is recommended pre-operatively, to manage hormonal imbalances caused by the NET, before its removal [361]. Caution is necessary in the case of insulinomas, where octreotide administration may even be fatal [283].
Apart from therapy per se, octreotide can be used for nuclear imaging purposes, in order to ascertain the SSTR expression of NETs, while it also enables the identification of possible occult tumors [362,363]; the most sensitive method available, making possible the detection of very small tumors is Positron Emission Tomography (PET) with gallium 68–labelled octreotide [364,365]. The sensitivity of PET in NETs is over 79% [366,367,368], except for insulinomas, where the 25% sensitivity [369] may be explained by the aforementioned diminished SSTR2 expression. The use of such sensitive methods is instrumental in the detection and staging associated with a more aggressive″ approach [361]. Such considerations aside, it was determined that in the case of imaging for neuroendocrine tumors, octreotide, and somatostatin analog treatments in general, do not seem to affect imaging in a negative way [370,371,372].

3.1.3. Octreotide in the Treatment of Lymphomas

Lymphomas are a diverse group of malignancies, generally arising from the clonal proliferation of lymphocytes [373,374]. The team of Dalm et al. [375] used octreotide in quantitative reverse transcription polymerase chain reaction autoradiography and determined that, in their study group, the expression of SSTR2 and SSTR3 seemed to be rather low.
The lower SSTR expression in lymphomas, at least compared to NETs, was corroborated by the subsequent research of Ferone et al. [376]. However, from a diagnostic point of view, the differential expression of SSTRs can be used to differentiate extragastric from gastric MALT-type lymphomas and is also a useful tool for staging the tumors and giving a prognosis [377]. On the other hand, there is a discussion regarding the higher specificity but lower sensitivity of octreotide scintigraphy, compared to more traditional methods of tumor imaging [378].
Apart from the aforementioned diagnostic value, studies mention that the expression of SSTRs may be used for treatment, in those lymphoma types that highly express SSTR2 [272]. But, for octreotide in particular, or indeed any somatostatin analog in general, to be useful therapeutically, the precise pattern of SSTR expression in lymphomas must be elucidated [205].

3.1.4. Octreotide in the Treatment of Merkel Cell Carcinoma

This is a rare, highly aggressive endocrine malignancy of the skin, with a mortality rate higher than that of melanoma [379,380,381]. As mentioned in the previous sections, Merkel carcinoma cells preferentially express SSTR2 and SSTR5 [188], meaning that in theory octreotide must be effective in such neoplasias. A treatment scheme of avelumab and octreotide, in a single patient, in a recent case report proved successful [274]. The subsequent study of Akaike et al. [275], who used somatostatin scintigraphy and then an octreotide-based treatment exhibited somewhat positive results; the promising potential of somatostatin analogs in Merkel cell carcinoma is also supported by the research of Anderson et al. who treated patients with monthly intramuscular injections of 20–30 mg with no significant adverse effects [189].

3.1.5. Octreotide in the Treatment of Hepatocellular Carcinoma (HCC)

Hepatic cancer is amongst the leading types of cancer at a worldwide level [382], with surprisingly high recurrence rates [383,384,385]; this is attributable to a variety of causes [386]. HCC, the most frequent hepatic cancer, is associated with a variety of risk factors [21,387,388]; its treatment or even management is exceedingly difficult, often requiring a complex approach [389,390,391]. Crucially for somatostatin analog treatment, hepatocellular carcinoma cells express all five SSTRs, which are not expressed in physiological hepatocytes; apart from HCC treatment, somatostatin analogs may also be used in cirrhotic liver patients [228,392], and have been successful in treating associated ascites [287,393,394].
Octreotide has been used extensively in the treatment of HCC, in numerous clinical trials. Although a number of such trials had a positive outcome [278,395,396,397,398,399,400,401,402,403,404,405,406,407,408,409], an equally substantial number of trials had a negative result [410,411,412,413,414,415,416,417]. The mixed results of the clinical trials may be due to the heterogeneity of the different studies; additionally, it must be stressed that many of the negative results are from trials involving end-stage patients [21]. Octreotide was administered in various forms including subcutaneously (150–750 μg/day) and intramuscular (30–40 mg/month).

3.2. Therapeutical Uses of Lanreotide in Neoplasias

Much like octrotide, lanreotide antagonizes hormone secretion and also has antiproliferative effects; it has high affinity for SSTR2 and SSTR5. Not so much data are available compared to octreotide, a fact consistent with the more limited use of lanreotide. The reported side effects are gastrointestinal, namely diarrhea, abdominal pain, nausea, vomiting, abdominal distention, impaired glucose tolerance, and gallstone formation, but also include cardiovascular side effects, alopecia, and pain and irritation at the injection site. Not all of these effects occur with the same frequencies and has not been established if and how much they are dose-dependent [268]. Lanreotide is not recommended during breastfeeding, although it is in case unlikely to reach toxic levels in the infant serum [418].
Just like octreotide, lanreotide can be used for the management of symptoms associated with NETs [273,419,420]. A clinical trial conducted by Caplin et al. [277] determined that the patients receiving lanreotide treatment (subcutaneous 120 mg/month aqueous-gel injection) had a better outcome, compared to placebo-receiving patients. Lanreotide autogel, is the approved initial treatment for patients with unresectable low-grade metastatic malignant insulinomas, effectively managing tumor growth and leading to hypoglycemia control in some but not all cases [276]. Also, a single case study [421] supports the use of lanreotide in duodenal neoplasm patients.
It must be noted that based on recent results, a half-year treatment with lanreotide in NET patients, did not significantly alter the patient’s perception of the disease [422]. However, NET patients seem to prefer lanreotide to octreotide in that it is easier to administer and is associated with fewer adverse reactions at the injection site [423]; another study also noted that satisfaction of patients with their lanreotide treatment regimen [424]. On the other hand, the randomized trial of Raj et al. [425] did not find significant differences between the preferences of NET patients for octreotide and lanreotide. Finally, in the treatment of hepatocellular carcinoma, the results of lanreotide administration were mixed, with one positive trial, in association with octreotide, [278] and one negative trial [410]; lanreotide is considered an overall safe drug [426,427] although adverse reactions have been cited, most recently after the administration for a pancreatic NET resulting in pituitary apoplexy [428].

3.3. Therapeutical Uses of Pasireotide in Neoplasias

This is a somatostatin analog activating most SSTRs, and having the lowest affinity for SSTR4. In addition to the hormones inhibited by octreotide and lanreotide, it also inhibits ACTH secretion [267]. Its side effects are very similar to those of lanreotide; while not likely to pass into breastmilk, lactating mothers are nonetheless advised to substitute it with an alternative treatment [429].
Pasireotide, being a newer drug, has not been extensively tested against NETs, but for a couple of clinical trials, such as Bruns et al. and Kvols et al. [281,430] which showed promising preliminary results. Given that desensitization to octreotide and lanreotide is a well-documented phenomenon [431]—possibly due to internalization or downregulation of SSTR2 in target cells, or overexpression of other SSTRs [432,433,434]—pasireotide, with its increased affinity for all SSTRs except for SSTR4, may offer a good alternative therapeutical approach [281]. It must be noted that perioperative treatment with pasireotide decreased the risk of significant pancreatic fistulas after pancreatic resection [435]. In general, the expression of SSTR2 by carcinoid tumors may allow for low doses of pasireotide, which, while still effective, will not be associated with significant side effects [436]. In another neoplasia, HCC, pasireotide was found to be ineffective [437,438].

4. Somatostatin Antagonists: Current Evidence and Studies

The role of somatostatin antagonists mainly revolves around radio imaging, based on the early concepts developed for single-photon emission computed tomography (SPECT) and PET [439]. Already SSTR imaging is a promising aspect in thyroid cancer [440]; SSTR-based imaging for neuroendocrine neoplasms is already an established approach [441], and generally, SSTR-based imaging can be used to monitor response to therapy and plan further therapeutic approaches [442]. Numerous radioligands have been developed that bind to the SSTR receptors providing essential anatomical and functional information through hybrid imaging, while also allowing for concurrent treatment via the so-called peptide receptor radionuclide therapy [443]. The precursors of most SSTR-targeting radiopharmaceuticals are the corresponding SSTR-targeting drugs already discussed; for example, the conversion of octreotide from agonist to antagonist is simple, involving just two positions [444].
Poorly internalized somatostatin antagonists are more effective for imaging, compared to highly internalized agonists [445]. Furthermore, it is believed that, at least in the case of some neuroendocrine tumors, the use of antagonist radiotracers may be more effective than the use of agonist radiotracers, as they can bind to a larger number of SSTR conformations [446]. These encouraging results were corroborated by the subsequent study of Fani et al. [447]. In treatment, the data on the use of SSTR antagonists are limited. There exist a few antagonists that have been tested for such purposes, with different affinities for some or all of SSTRs.
One characteristic case is the research of Modarai et al. [448] on colorectal cancer; they hypothesized that SSTR1 was instrumental in maintaining the quiescence of colonic stem cells; their abnormal proliferation leads to colorectal cancer. In an in vitro study, they determined that somatostatin was not expressed in cancer cells, in contrast with SSTR1 which was expressed. Regardless, the precise role significance of certain SSTRs in colorectal cancer, when correlated with the local microenvironment and microbiome has not yet been elucidated [449]. The use of cyclosomatostatin was important in differentiating between the different mechanisms controlling the multiplication and differentiation of stem cells [448]. Up to this point, cyclosomatostatin has been used in a number of other animal experiments [450,451,452,453,454,455,456,457,458,459,460]; to our knowledge, no clinical use of it has yet been made.
Recent studies explore novel compounds with therapeutic aims, and their area of relevance might be expanded to cancer treatment [461,462,463]. Following up on the results of Sprecher et al. [464], the team of Hirose et al. [465] developed and tested a 5-oxa-2,6-diazaspiro [3.4]oct-6-ene derivative, which is a novel SSTR5 antagonist, effective against type 2 diabetes mellitus in mice, as it augmented glucose-dependent insulin secretion. Another antagonist, CYN 154806, a solid cyclic octapeptide, has been successfully tested in vitro [466] and in vivo in mice [467]. The SSTR3-selective MK-4256, has been successfully tested and is able to reduce glucose levels in mice [468]. PRL 2915, with high affinity for SSTR3 and a lower affinity for SSTR5 [469,470], has been successfully applied in vitro to stop aortic ring contractions [470]. BIM-23056, a linear octapeptide with SSTR affinities similar to that of the previous antagonist, has not been very successful in vitro, as of yet [471]. Finally, another novel somatostatin antagonist, a spiropiperidine analog, proposed by Liu et al. [472] may prove useful in the management of diabetes mellitus, if optimized to achieve the necessary pharmacokinetic profile.

5. Discussion

SSTRs have numerous important physiological functions, which depend on their associated signaling mechanisms and distribution in the human body. The differential distribution of SSTRs in cancers enables the selective use of somatostatin analogs, as viable solutions in the treatment of neoplasias or associated symptoms, with reasonably few side effects, in many cases. From the three analogs currently in clinical use, octreotide, which has been in constant use for over four decades now, is characterized by ample clinical data both regarding its effectiveness and side effects; for the other agonists and antagonists, evidence is encouraging but fewer in comparison.
Apart from the cases of somatostatin agonists and antagonists presented above, there are also other compounds, whose application may provide new insights. Such an example is dopastatin, a chemical compound combining the structure of dopamine and somatostatin; the use of this compound was envisaged as a therapeutic agent in patients with acromegaly, and indeed, its ability to act both on SSTR2 and dopamine D2 receptors was found to greatly enhance its effects [473,474]. While this enhanced effect has not been yet explained, Rocheville et al. [475] note that it is possible to attribute it to a heterodimerization between the two aforementioned receptors. Based on these data, it is possible that, in the future, chimeric compounds combining somatostatin with other compatible hormones, could be used, especially in the case of multi-hormone-dependent tumors, such as breast cancer and prostate cancer [476] and ovarian cancer [477,478]. Combination therapy of somatostatin agonists and various chemotherapeutic drugs is being researched for a variety of cancers, including NETs, with some favorable results regarding response rates and progression-free survival [479].
The use of antagonists in combination with radioisotopes, for cancer therapy is another opportunity, given the high efficiency of antagonists in binding to specific receptors, as demonstrated in the relevant imaging studies. Such a type of therapy was the subject of the research of Kong et al. [480], who achieved good results in patients with bulky neuroendocrine tumors. In cancers, the main application of agonists is to inhibit hormone secretion and exhibit antiproliferative effects through cell cycle arrest and pro-apoptotic pathways; conversely, antagonists may demonstrate antitumoral effects by increasing the secretion of certain growth-inhibiting factors and pro-apoptotic cytokines as well as through the disruption of the tumoral adaptive mechanisms with secondary restoration of tumor sensitivity to the treatment [481].
Based on the aforementioned research results, and regarding the distribution of SSTRs in different pathologies, it can be seen that in some neoplasias, there is differential SSTR expression which may be important for staging, prognosis and therapy. Nevertheless, there are some cases where, for the same type of cancer, there are conflicting research results. As such, there are still certain cases where the presence of SSTRs and their relative expression has not been definitely assessed as to their significance and the associated therapeutic potential.
The study of Angelousi et al. [482], focused on the resistance patterns of NETs to somatostatin analog treatment, highlighting the considerable diversity of tumor resistance, which is presumably associated with the particular molecular and genetic mechanisms of each different tumor. Such data are indicative of the need to personalize treatments based on differential tumor parameters and responses. Other important issues in somatostatin analog treatment are the determination of the most effective regimens as well as whether dose-dependent effects occur [483,484,485].
Apart from the neoplastic syndromes mentioned in our analysis, it is possible that other neoplasias may be characterized by SSTR expression; this could be the focus of future research. Moreover, and taking into account the known resistance patterns to classic somatostatin analogs and their side effects, the introduction of other, non-peptide, SSTR agonists, such as the compound used by Juliana et al. [486], may improve the current treatment schemes. The treatment of neoplasia patients is oftentimes a complicated issue [487], as seen in the case of NET patients, and, especially when it includes somatostatin analogs, should be performed at dedicated and experienced centers [488]. Moreover, numerous alterations within the tumor microenvironment may be reflected by serum changes of various metabolites or cell expression, leading to further interferences between carcinogenesis and the immune response [449,489,490,491].
Interestingly, despite the known side effects of somatostatin analogs, an increased percentage of acromegaly and NET patients are still given the maximum allowed analog dose, at least in some regions [492]. It has also been found, in acromegaly patients, that prolonged octreotide treatment does not seem to increase Helicobacter pylori incidence [493] and despite earlier ambiguous research results [494], it was even recently demonstrated, in rats, that octreotide treatment can protect against H. pylori-induced gastritis [495]. Although there is no direct link between acromegaly and gastritis, it was observed that octreotide treatment was associated with increased gastritis incidence, and it was believed that H. pylori might represent a causative factor [493,494].
Apart from the administration of somatostatin analogs, another avenue that should be considered is the induction of endogenous somatostatin production. It has been demonstrated that systemic capsaicin administration can cause the elevation of somatostatin levels [496,497,498,499]. This might hold therapeutical potential, perhaps by local instillation of capsaicin, thus obviating the risk of capsaicin side effects that accompany its systemic administration in high doses [500,501,502,503,504,505]. The interplay with SSTRs and somatostatin levels in cancers is of interest since capsaicin itself also exhibits antitumoral effects [506,507,508,509]. Another possibility is the use of nanoparticles, which are already been studied for the deployment of anti-cancer agents [510,511,512,513,514], or lipoparticles [515,516] for targeted delivery to neoplasias.
Modulation of endogenous somatostatin production might be achieved through acupuncture, and literature data showcase the effects of acupuncture on serum hormone and protein levels both in animal models and humans [517,518,519,520,521,522,523]. Moreover, acupuncture has also been shown to be effective in the management of cancer-related symptoms [524,525,526,527]. Specifically for somatostatin secretion induction, in rat and dog models, electroacupuncture has been demonstrated to increase serum somatostatin levels [528,529,530,531,532,533,534,535] and possibly SSTR expression in rabbits [536]. In humans, electroacupuncture and moxibustion could also influence serum somatostatin levels [537,538] and, in turn, it has been proposed that the endogenous levels of somatostatin may influence the effects of electroacupuncture [539,540].
Finally, as analyzed above, it is possible to use somatostatin agonists or antagonists, to improve imaging methods in certain cases. Another possibility nowadays is imaging-controlled biopsies such as Magnetic Resonance Imaging (MRI)-transrectal ultrasound fusion guided prostate biopsy [541,542]; maybe it is also possible to consider the use of radiolabelled agonists or antagonists for PET/Computed Tomography (CT)-guided biopsies, which have already entered clinical practice [543,544,545]. For a number of pathologies, like thymoma, MRI is considered to be the best imaging method [546,547], but perhaps the sensitivity of PET/CT could be augmented by using radiolabelled somatostatin analogs. Additionally, fluorescent somatostatin analogs were used in order to delineate tumor boundaries for SSTR2-expressing tumors during surgery [548]. This opens new research avenues for enhancing imaging techniques with clinical applications in oncology, by improving pre- and intraoperative as well as follow-up diagnostic performance [549,550,551]; these applications might also allow for improvements in robotic surgery [552,553]. Future enhancements may also include the combination of SSTR imaging with 3D printing in order to aid in tumor identification and separation [554,555,556,557].
There are several challenges to translating basic research on SSTRs into clinical practice. Firstly, the complexity of SSTR expression in various tumor types and between individuals can lead to inconsistent therapeutic responses [558,559]. Additionally, overcoming drug resistance caused by receptor desensitization or tumor adaptation needs further development of therapeutic agents or antagonists [560]. Lastly, translating results from preclinical animal studies to human patients may be hindered by the species-specific structure and function of SSTRs, leading to the necessity of significant phase II and III studies in order to establish therapeutic safety and efficiency [479].

6. Conclusions

The gradual introduction of SSTR antagonists into clinical trial status may reveal new intervention options. Despite a fair amount of research on the pathophysiological associations of somatostatin and its related drugs, open questions yet remain, which are interesting questions for further research.
While the interaction of somatostatin with its receptors is well documented, it remains to be seen if it, or its analogs, are able to interact with other extracellular or intracellular receptors and if such an interaction is associated with different effects. In that regard, the synthesis and use of chimeric somatostatin molecules seem an interesting research avenue. The existence of receptors sensitive to somatostatin and its analogs, other than the known SSTRs, can be explored in the context of therapeutic resistance exhibited by some cancers when treated with somatostatin analogs.
On the other hand, somatostatin antagonists, which are coming to the foreground of relevant medical research, should be examined for their potential to avert or mitigate the side effects of somatostatin agonists, without compromising their therapeutical potential.
Apart from all these considerations and intriguing research perspectives, the continued research on the properties and structure of SSTRs will, hopefully, facilitate the discovery of new, even more selective analogs, with higher therapeutic potential.

Author Contributions

Conceptualization, A.P., G.T., C.S. and C.C.; formal analysis, C.M.T., D.A.H., I.A.B., C.S. and A.C.; investigation, A.P., G.T., A.-T.P. and L.T.; resources, K.P., I.S.-F., C.C. and A.-E.S.; data curation, L.T.; writing—original draft preparation, A.P., G.T., A.-T.P., C.M.T., D.A.H., K.P., L.T., I.A.B., C.S., A.C., I.S.-F., C.C. and A.-E.S.; writing—review and editing, A.P., C.S., C.C. and A.-E.S.; supervision, C.S., I.S.-F., C.C. and A.-E.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Brazeau, P.; Vale, W.; Burgus, R.; Ling, N.; Butcher, M.; Rivier, J.; Guillemin, R. Hypothalamic polypeptide that inhibits the secretion of immunoreactive pituitary growth hormone. Science 1973, 179, 77–79. [Google Scholar] [CrossRef] [PubMed]
  2. Esch, F.; Böhlen, P.; Ling, N.; Benoit, R.; Brazeau, P.; Guillemin, R. Primary structure of ovine hypothalamic somatostatin-28 and somatostatin-25. Proc. Natl. Acad. Sci. USA 1980, 77, 6827–6831. [Google Scholar] [CrossRef]
  3. Pradayrol, L.; Jörnvall, H.; Mutt, V.; Ribet, A. N-terminally extended somatostatin: The primary structure of somatostatin-28. FEBS Lett. 1980, 109, 55–58. [Google Scholar] [CrossRef]
  4. Shen, L.P.; Rutter, W.J. Sequence of the human somatostatin I gene. Science 1984, 224, 168–171. [Google Scholar] [CrossRef] [PubMed]
  5. Günther, T.; Tulipano, G.; Dournaud, P.; Bousquet, C.; Csaba, Z.; Kreienkamp, H.J.; Lupp, A.; Korbonits, M.; Castaño, J.P.; Wester, H.J.; et al. International Union of Basic and Clinical Pharmacology. CV. Somatostatin Receptors: Structure, Function, Ligands, and New Nomenclature. Pharmacol. Rev. 2018, 70, 763–835. [Google Scholar] [CrossRef] [PubMed]
  6. Reichlin, S. Somatostatin. N. Engl. J. Med. 1983, 309, 1495–1501. [Google Scholar] [CrossRef]
  7. Reichlin, S. Somatostatin (second of two parts). N. Engl. J. Med. 1983, 309, 1556–1563. [Google Scholar] [CrossRef]
  8. Hall, R.; Page, M.D.; Dieguez, C.; Scanlon, M.F. Somatostatin: A Historical Perspective. Horm. Res. 2008, 29, 50–53. [Google Scholar] [CrossRef]
  9. Reisine, T.; Bell, G.I. Molecular biology of somatostatin receptors. Endocr. Rev. 1995, 16, 427–442. [Google Scholar] [CrossRef]
  10. O’Toole, T.J.; Sharma, S. Physiology, Somatostatin. In StatPearls; StatPearls Publishing LLC: Treasure Island, FL, USA, 2024. [Google Scholar]
  11. Bloom, S.R.; Polak, J.M. Somatostatin. Br. Med. J. 1987, 295, 288–290. [Google Scholar] [CrossRef]
  12. Cervia, D.; Casini, G.; Bagnoli, P. Physiology and pathology of somatostatin in the mammalian retina: A current view. Mol. Cell. Endocrinol. 2008, 286, 112–122. [Google Scholar] [CrossRef] [PubMed]
  13. Scheau, C.; Draghici, C.; Ilie, M.A.; Lupu, M.; Solomon, I.; Tampa, M.; Georgescu, S.R.; Caruntu, A.; Constantin, C.; Neagu, M.; et al. Neuroendocrine Factors in Melanoma Pathogenesis. Cancers 2021, 13, 2277. [Google Scholar] [CrossRef] [PubMed]
  14. Harris, A.G. Somatostatin and somatostatin analogues: Pharmacokinetics and pharmacodynamic effects. Gut 1994, 35, S1–S4. [Google Scholar] [CrossRef]
  15. Theodoropoulou, M.; Stalla, G.K. Somatostatin receptors: From signaling to clinical practice. Front. Neuroendocrinol. 2013, 34, 228–252. [Google Scholar] [CrossRef]
  16. Shamsi, B.H.; Chatoo, M.; Xu, X.K.; Xu, X.; Chen, X.Q. Versatile Functions of Somatostatin and Somatostatin Receptors in the Gastrointestinal System. Front. Endocrinol. 2021, 12, 652363. [Google Scholar] [CrossRef]
  17. Robinson, S.L.; Thiele, T.E. Somatostatin signaling modulates binge drinking behavior via the central nucleus of the amygdala. Neuropharmacology 2023, 237, 109622. [Google Scholar] [CrossRef]
  18. Joye, D.A.M.; Rohr, K.E.; Suenkens, K.; Wuorinen, A.; Inda, T.; Arzbecker, M.; Mueller, E.; Huber, A.; Pancholi, H.; Blackmore, M.G.; et al. Somatostatin regulates central clock function and circadian responses to light. Proc. Natl. Acad. Sci. USA 2023, 120, e2216820120. [Google Scholar] [CrossRef] [PubMed]
  19. Schwartz, J.P.; Ji, Z.; Epelbaum, J. Somatostatin as a neurotrophic factor. Which receptor/second messenger transduction system is involved? Perspect. Dev. Neurobiol. 1998, 5, 427–435. [Google Scholar]
  20. Rai, U.; Thrimawithana, T.R.; Valery, C.; Young, S.A. Therapeutic uses of somatostatin and its analogues: Current view and potential applications. Pharmacol. Ther. 2015, 152, 98–110. [Google Scholar] [CrossRef]
  21. Periferakis, A.; Tsigas, G.; Periferakis, A.T.; Badarau, I.A.; Scheau, A.E.; Tampa, M.; Georgescu, S.R.; Didilescu, A.C.; Scheau, C.; Caruntu, C. Antitumoral and Anti-inflammatory Roles of Somatostatin and Its Analogs in Hepatocellular Carcinoma. Anal. Cell. Pathol. 2021, 2021, 1840069. [Google Scholar] [CrossRef]
  22. Kouroumalis, E.; Samonakis, D.; Notas, G. Somatostatin in hepatocellular carcinoma: Experimental and therapeutic implications. Hepatoma Res. 2018, 4, 34. [Google Scholar] [CrossRef]
  23. Liu, M.; Cheng, Y.; Bai, C.; Zhao, H.; Jia, R.; Chen, J.; Zhu, W.; Huo, L. Gallium-68 labeled somatostatin receptor antagonist PET/CT in over 500 patients with neuroendocrine neoplasms: Experience from a single center in China. Eur. J. Nucl. Med. Mol. Imaging 2024, 51, 2002–2011. [Google Scholar] [CrossRef] [PubMed]
  24. Han, F.; Zhao, T.; Zhang, Y.; Yun, Y.; Xu, Y.; Guo, S.; Zhong, Y.; Xie, X.; Shen, J. Discovery and exploration of novel somatostatin receptor subtype 5 (SSTR5) antagonists for the treatment of cholesterol gallstones. Eur. J. Med. Chem. 2024, 264, 116017. [Google Scholar] [CrossRef]
  25. Vanhoutte, P.M.; Humphrey, P.P.; Spedding, M.X. International Union of Pharmacology recommendations for nomenclature of new receptor subtypes. Pharmacol. Rev. 1996, 48, 1–2. [Google Scholar]
  26. Hoyer, D.; Bell, G.I.; Berelowitz, M.; Epelbaum, J.; Feniuk, W.; Humphrey, P.P.; O’Carroll, A.M.; Patel, Y.C.; Schonbrunn, A.; Taylor, J.E.; et al. Classification and nomenclature of somatostatin receptors. Trends Pharmacol. Sci. 1995, 16, 86–88. [Google Scholar] [CrossRef]
  27. Yamada, Y.; Post, S.R.; Wang, K.; Tager, H.S.; Bell, G.I.; Seino, S. Cloning and functional characterization of a family of human and mouse somatostatin receptors expressed in brain, gastrointestinal tract, and kidney. Proc. Natl. Acad. Sci. USA 1992, 89, 251–255. [Google Scholar] [CrossRef] [PubMed]
  28. Schulz, S.; Händel, M.; Schreff, M.; Schmidt, H.; Höllt, V. Localization of five somatostatin receptors in the rat central nervous system using subtype-specific antibodies. J. Physiol. 2000, 94, 259–264. [Google Scholar] [CrossRef]
  29. Redmann, A.; Rasch, A.; Tourné, H.; Mann, K.; Petersenn, S. Characterization and transcriptional regulation of the human somatostatin receptor subtype 1 gene. Horm. Metab. Res. 2007, 39, 359–365. [Google Scholar] [CrossRef]
  30. Van Op den Bosch, J.; van Nassauw, L.; Lantermann, K.; van Marck, E.; Timmermans, J.P. Effect of intestinal inflammation on the cell-specific expression of somatostatin receptor subtypes in the murine ileum. Neurogastroenterol. Motil. 2007, 19, 596–606. [Google Scholar] [CrossRef]
  31. Lupp, A.; Nagel, F.; Schulz, S. Reevaluation of sst1 somatostatin receptor expression in human normal and neoplastic tissues using the novel rabbit monoclonal antibody UMB-7. Regul. Pept. 2013, 183, 1–6. [Google Scholar] [CrossRef]
  32. Rohrer, L.; Raulf, F.; Bruns, C.; Buettner, R.; Hofstaedter, F.; Schüle, R. Cloning and characterization of a fourth human somatostatin receptor. Proc. Natl. Acad. Sci. USA 1993, 90, 4196–4200. [Google Scholar] [CrossRef]
  33. Fischer, T.; Doll, C.; Jacobs, S.; Kolodziej, A.; Stumm, R.; Schulz, S. Reassessment of sst2 somatostatin receptor expression in human normal and neoplastic tissues using the novel rabbit monoclonal antibody UMB-1. J. Clin. Endocrinol. Metab. 2008, 93, 4519–4524. [Google Scholar] [CrossRef]
  34. Chin, S.O.; Ku, C.R.; Kim, B.J.; Kim, S.W.; Park, K.H.; Song, K.H.; Oh, S.; Yoon, H.K.; Lee, E.J.; Lee, J.M.; et al. Medical Treatment with Somatostatin Analogues in Acromegaly: Position Statement. Endocrinol. Metab. 2019, 34, 53–62. [Google Scholar] [CrossRef] [PubMed]
  35. Yamada, Y.; Reisine, T.; Law, S.F.; Ihara, Y.; Kubota, A.; Kagimoto, S.; Seino, M.; Seino, Y.; Bell, G.I.; Seino, S. Somatostatin receptors, an expanding gene family: Cloning and functional characterization of human SSTR3, a protein coupled to adenylyl cyclase. Mol. Endocrinol. 1992, 6, 2136–2142. [Google Scholar] [CrossRef] [PubMed]
  36. Lupp, A.; Nagel, F.; Doll, C.; Röcken, C.; Evert, M.; Mawrin, C.; Saeger, W.; Schulz, S. Reassessment of sst3 somatostatin receptor expression in human normal and neoplastic tissues using the novel rabbit monoclonal antibody UMB-5. Neuroendocrinology 2012, 96, 301–310. [Google Scholar] [CrossRef] [PubMed]
  37. Demchyshyn, L.L.; Srikant, C.B.; Sunahara, R.K.; Kent, G.; Seeman, P.; Van Tol, H.H.; Panetta, R.; Patel, Y.C.; Niznik, H.B. Cloning and expression of a human somatostatin-14-selective receptor variant (somatostatin receptor 4) located on chromosome 20. Mol. Pharmacol. 1993, 43, 894–901. [Google Scholar]
  38. Schreff, M.; Schulz, S.; Händel, M.; Keilhoff, G.; Braun, H.; Pereira, G.; Klutzny, M.; Schmidt, H.; Wolf, G.; Höllt, V. Distribution, targeting, and internalization of the sst4 somatostatin receptor in rat brain. J. Neurosci. 2000, 20, 3785–3797. [Google Scholar] [CrossRef] [PubMed]
  39. Bär, K.J.; Schurigt, U.; Scholze, A.; Segond Von Banchet, G.; Stopfel, N.; Bräuer, R.; Halbhuber, K.J.; Schaible, H.G. The expression and localization of somatostatin receptors in dorsal root ganglion neurons of normal and monoarthritic rats. Neuroscience 2004, 127, 197–206. [Google Scholar] [CrossRef]
  40. Taniyama, Y.; Suzuki, T.; Mikami, Y.; Moriya, T.; Satomi, S.; Sasano, H. Systemic distribution of somatostatin receptor subtypes in human: An immunohistochemical study. Endocr. J. 2005, 52, 605–611. [Google Scholar] [CrossRef]
  41. Panetta, R.; Greenwood, M.T.; Warszynska, A.; Demchyshyn, L.L.; Day, R.; Niznik, H.B.; Srikant, C.B.; Patel, Y.C. Molecular cloning, functional characterization, and chromosomal localization of a human somatostatin receptor (somatostatin receptor type 5) with preferential affinity for somatostatin-28. Mol. Pharmacol. 1994, 45, 417–427. [Google Scholar]
  42. Takeda, J.; Fernald, A.A.; Yamagata, K.; Le Beau, M.M.; Bell, G.I. Localization of human somatostatin receptor 5 gene (SSTR5) to chromosome band 16p13.3 by fluorescence in situ hybridization. Genomics 1995, 26, 638–639. [Google Scholar] [CrossRef] [PubMed]
  43. Stollberg, S.; Kämmerer, D.; Neubauer, E.; Schulz, S.; Simonitsch-Klupp, I.; Kiesewetter, B.; Raderer, M.; Lupp, A. Differential somatostatin and CXCR4 chemokine receptor expression in MALT-type lymphoma of gastric and extragastric origin. J. Cancer Res. Clin. Oncol. 2016, 142, 2239–2247. [Google Scholar] [CrossRef] [PubMed]
  44. Gadelha, M.R.; Wildemberg, L.E.; Bronstein, M.D.; Gatto, F.; Ferone, D. Somatostatin receptor ligands in the treatment of acromegaly. Pituitary 2017, 20, 100–108. [Google Scholar] [CrossRef] [PubMed]
  45. Pintér, E.; Helyes, Z.; Szolcsányi, J. Inhibitory effect of somatostatin on inflammation and nociception. Pharmacol. Ther. 2006, 112, 440–456. [Google Scholar] [CrossRef] [PubMed]
  46. Casini, G.; Catalani, E.; Dal Monte, M.; Bagnoli, P. Functional aspects of the somatostatinergic system in the retina and the potential therapeutic role of somatostatin in retinal disease. Histol. Histopathol. 2005, 20, 615–632. [Google Scholar] [CrossRef]
  47. Engin, E.; Stellbrink, J.; Treit, D.; Dickson, C.T. Anxiolytic and antidepressant effects of intracerebroventricularly administered somatostatin: Behavioral and neurophysiological evidence. Neuroscience 2008, 157, 666–676. [Google Scholar] [CrossRef]
  48. Engin, E.; Treit, D. Anxiolytic and antidepressant actions of somatostatin: The role of sst2 and sst3 receptors. Psychopharmacology 2009, 206, 281–289. [Google Scholar] [CrossRef]
  49. Vasilaki, A.; Thermos, K. Somatostatin analogues as therapeutics in retinal disease. Pharmacol. Ther. 2009, 122, 324–333. [Google Scholar] [CrossRef]
  50. Ben-Shlomo, A.; Melmed, S. Pituitary somatostatin receptor signaling. Trends Endocrinol. Metab. 2010, 21, 123–133. [Google Scholar] [CrossRef]
  51. Stengel, A.; Karasawa, H.; Taché, Y. The role of brain somatostatin receptor 2 in the regulation of feeding and drinking behavior. Horm. Behav. 2015, 73, 15–22. [Google Scholar] [CrossRef]
  52. Prévôt, T.D.; Gastambide, F.; Viollet, C.; Henkous, N.; Martel, G.; Epelbaum, J.; Béracochéa, D.; Guillou, J.L. Roles of Hippocampal Somatostatin Receptor Subtypes in Stress Response and Emotionality. Neuropsychopharmacology 2017, 42, 1647–1656. [Google Scholar] [CrossRef] [PubMed]
  53. Patel, Y.C. Somatostatin and its receptor family. Front. Neuroendocrinol. 1999, 20, 157–198. [Google Scholar] [CrossRef] [PubMed]
  54. Mergler, S.; Singh, V.; Grötzinger, C.; Kaczmarek, P.; Wiedenmann, B.; Strowski, M.Z. Characterization of voltage operated R-type Ca2+ channels in modulating somatostatin receptor subtype 2- and 3-dependent inhibition of insulin secretion from INS-1 cells. Cell. Signal. 2008, 20, 2286–2295. [Google Scholar] [CrossRef] [PubMed]
  55. Eigler, T.; Ben-Shlomo, A.; Zhou, C.; Khalafi, R.; Ren, S.G.; Melmed, S. Constitutive somatostatin receptor subtype-3 signaling suppresses growth hormone synthesis. Mol. Endocrinol. 2014, 28, 554–564. [Google Scholar] [CrossRef]
  56. Yao, Q.; Liu, Q.; Xu, H.; Wu, Z.; Zhou, L.; Gu, Z.; Gong, P.; Shen, J. Upregulated Expression of SSTR3 is Involved in Neuronal Apoptosis after Intracerebral Hemorrhage in Adult Rats. Cell Mol. Neurobiol. 2017, 37, 1407–1416. [Google Scholar] [CrossRef]
  57. Santis, S.; Kastellakis, A.; Kotzamani, D.; Pitarokoili, K.; Kokona, D.; Thermos, K. Somatostatin increases rat locomotor activity by activating sst2 and sst4 receptors in the striatum and via glutamatergic involvement. Naunyn Schmiedebergs Arch. Pharmacol. 2009, 379, 181–189. [Google Scholar] [CrossRef] [PubMed]
  58. Gastambide, F.; Lepousez, G.; Viollet, C.; Loudes, C.; Epelbaum, J.; Guillou, J.L. Cooperation between hippocampal somatostatin receptor subtypes 4 and 2: Functional relevance in interactive memory systems. Hippocampus 2010, 20, 745–757. [Google Scholar] [CrossRef]
  59. Martínez, V.; Rivier, J.; Coy, D.; Taché, Y. Intracisternal injection of somatostatin receptor 5-preferring agonists induces a vagal cholinergic stimulation of gastric emptying in rats. J. Pharmacol. Exp. Ther. 2000, 293, 1099–1105. [Google Scholar]
  60. Stengel, A.; Taché, Y.F. Activation of Brain Somatostatin Signaling Suppresses CRF Receptor-Mediated Stress Response. Front. Neurosci. 2017, 11, 231. [Google Scholar] [CrossRef]
  61. Costanzi, S.; Siegel, J.; Tikhonova, I.G.; Jacobson, K.A. Rhodopsin and the others: A historical perspective on structural studies of G protein-coupled receptors. Curr. Pharm. Des. 2009, 15, 3994–4002. [Google Scholar] [CrossRef]
  62. Kobilka, B.K. G protein coupled receptor structure and activation. Biochim. Biophys. Acta (BBA)—Biomembr. 2007, 1768, 794–807. [Google Scholar] [CrossRef] [PubMed]
  63. Zhang, B.; Xue, L.; Wu, Z.B. Structure and Function of Somatostatin and its Receptors in Endocrinology. Endocr. Rev. 2024, bnae022. [Google Scholar] [CrossRef] [PubMed]
  64. Vanetti, M.; Kouba, M.; Wang, X.; Vogt, G.; Höllt, V. Cloning and expression of a novel mouse somatostatin receptor (SSTR2B). FEBS Lett. 1992, 311, 290–294. [Google Scholar] [CrossRef]
  65. Durán-Prado, M.; Gahete, M.D.; Martínez-Fuentes, A.J.; Luque, R.M.; Quintero, A.; Webb, S.M.; Benito-López, P.; Leal, A.; Schulz, S.; Gracia-Navarro, F.; et al. Identification and characterization of two novel truncated but functional isoforms of the somatostatin receptor subtype 5 differentially present in pituitary tumors. J. Clin. Endocrinol. Metab. 2009, 94, 2634–2643. [Google Scholar] [CrossRef] [PubMed]
  66. Greenwood, M.T.; Robertson, L.A.; Patel, Y.C. Cloning of the gene encoding human somatostatin receptor 2: Sequence analysis of the 5′-flanking promoter region. Gene 1995, 159, 291–292. [Google Scholar] [CrossRef]
  67. Kiran, K.; Ansari, S.A.; Srivastava, R.; Lodhi, N.; Chaturvedi, C.P.; Sawant, S.V.; Tuli, R. The TATA-box sequence in the basal promoter contributes to determining light-dependent gene expression in plants. Plant Physiol. 2006, 142, 364–376. [Google Scholar] [CrossRef]
  68. Wu, W.-S.; Lai, F.-J. Functional redundancy of transcription factors explains why most binding targets of a transcription factor are not affected when the transcription factor is knocked out. BMC Syst. Biol. 2015, 9, S2. [Google Scholar] [CrossRef]
  69. Carruthers, A.M.; Warner, A.J.; Michel, A.D.; Feniuk, W.; Humphrey, P.P. Activation of adenylate cyclase by human recombinant sst5 receptors expressed in CHO-K1 cells and involvement of Galphas proteins. Br. J. Pharmacol. 1999, 126, 1221–1229. [Google Scholar] [CrossRef]
  70. Gromada, J.; Høy, M.; Buschard, K.; Salehi, A.; Rorsman, P. Somatostatin inhibits exocytosis in rat pancreatic alpha-cells by Gi2-dependent activation of calcineurin and depriming of secretory granules. J. Physiol. 2001, 535, 519–532. [Google Scholar] [CrossRef]
  71. Pan, M.G.; Florio, T.; Stork, P.J. G protein activation of a hormone-stimulated phosphatase in human tumor cells. Science 1992, 256, 1215–1217. [Google Scholar] [CrossRef]
  72. Csaba, Z.; Dournaud, P. Cellular biology of somatostatin receptors. Neuropeptides 2001, 35, 1–23. [Google Scholar] [CrossRef] [PubMed]
  73. Durán-Prado, M.; Malagón, M.M.; Gracia-Navarro, F.; Castaño, J.P. Dimerization of G protein-coupled receptors: New avenues for somatostatin receptor signalling, control and functioning. Mol. Cell. Endocrinol. 2008, 286, 63–68. [Google Scholar] [CrossRef] [PubMed]
  74. Böhm, S.K.; Grady, E.F.; Bunnett, N.W. Regulatory mechanisms that modulate signalling by G-protein-coupled receptors. Biochem. J. 1997, 322 Pt 1, 1–18. [Google Scholar] [CrossRef]
  75. Christofides, K.; Menon, R.; Jones, C.E. Endocytosis of G Protein-Coupled Receptors and Their Ligands: Is There a Role in Metal Trafficking? Cell Biochem. Biophys. 2018, 76, 329–337. [Google Scholar] [CrossRef]
  76. Fani, M.; Nicolas, G.P.; Wild, D. Somatostatin Receptor Antagonists for Imaging and Therapy. J. Nucl. Med. 2017, 58, 61s–66s. [Google Scholar] [CrossRef] [PubMed]
  77. Csaba, Z.; Dournaud, P. Internalization of somatostatin receptors in brain and periphery. Progress Mol. Biol. Transl. Sci. 2023, 196, 43–57. [Google Scholar] [CrossRef]
  78. Vasilaki, A.; Papasava, D.; Hoyer, D.; Thermos, K. The somatostatin receptor (sst1) modulates the release of somatostatin in the nucleus accumbens of the rat. Neuropharmacology 2004, 47, 612–618. [Google Scholar] [CrossRef]
  79. Tannenbaum, G.S.; Zhang, W.H.; Lapointe, M.; Zeitler, P.; Beaudet, A. Growth hormone-releasing hormone neurons in the arcuate nucleus express both Sst1 and Sst2 somatostatin receptor genes. Endocrinology 1998, 139, 1450–1453. [Google Scholar] [CrossRef]
  80. Lanneau, C.; Bluet-Pajot, M.T.; Zizzari, P.; Csaba, Z.; Dournaud, P.; Helboe, L.; Hoyer, D.; Pellegrini, E.; Tannenbaum, G.S.; Epelbaum, J.; et al. Involvement of the Sst1 somatostatin receptor subtype in the intrahypothalamic neuronal network regulating growth hormone secretion: An in vitro and in vivo antisense study. Endocrinology 2000, 141, 967–979. [Google Scholar] [CrossRef]
  81. Thermos, K.; Bagnoli, P.; Epelbaum, J.; Hoyer, D. The somatostatin sst1 receptor: An autoreceptor for somatostatin in brain and retina? Pharmacol. Ther. 2006, 110, 455–464. [Google Scholar] [CrossRef]
  82. Momiyama, T.; Zaborszky, L. Somatostatin presynaptically inhibits both GABA and glutamate release onto rat basal forebrain cholinergic neurons. J. Neurophysiol. 2006, 96, 686–694. [Google Scholar] [CrossRef] [PubMed]
  83. Malcangio, M. GDNF and somatostatin in sensory neurones. Curr. Opin. Pharmacol. 2003, 3, 41–45. [Google Scholar] [CrossRef] [PubMed]
  84. Imhof, A.K.; Glück, L.; Gajda, M.; Lupp, A.; Bräuer, R.; Schaible, H.G.; Schulz, S. Differential antiinflammatory and antinociceptive effects of the somatostatin analogs octreotide and pasireotide in a mouse model of immune-mediated arthritis. Arthritis Rheum. 2011, 63, 2352–2362. [Google Scholar] [CrossRef]
  85. Portela-Gomes, G.M.; Stridsberg, M.; Grimelius, L.; Oberg, K.; Janson, E.T. Expression of the five different somatostatin receptor subtypes in endocrine cells of the pancreas. Appl. Immunohistochem. Mol. Morphol. 2000, 8, 126–132. [Google Scholar] [CrossRef] [PubMed]
  86. Wang, X.P.; Norman, M.A.; Yang, J.; Cheung, A.; Moldovan, S.; Demayo, F.J.; Brunicardi, F.C. Double-gene ablation of SSTR1 and SSTR5 results in hyperinsulinemia and improved glucose tolerance in mice. Surgery 2004, 136, 585–592. [Google Scholar] [CrossRef] [PubMed]
  87. Smith, P.A. N-type Ca2+-channels in murine pancreatic beta-cells are inhibited by an exclusive coupling with somatostatin receptor subtype 1. Endocrinology 2009, 150, 741–748. [Google Scholar] [CrossRef]
  88. Garcia, P.D.; Myers, R.M. Pituitary cell line GH3 expresses two somatostatin receptor subtypes that inhibit adenylyl cyclase: Functional expression of rat somatostatin receptor subtypes 1 and 2 in human embryonic kidney 293 cells. Mol. Pharmacol. 1994, 45, 402–409. [Google Scholar]
  89. Hadcock, J.R.; Strnad, J.; Eppler, C.M. Rat somatostatin receptor type 1 couples to G proteins and inhibition of cyclic AMP accumulation. Mol. Pharmacol. 1994, 45, 410–416. [Google Scholar]
  90. Hershberger, R.E.; Newman, B.L.; Florio, T.; Bunzow, J.; Civelli, O.; Li, X.J.; Forte, M.; Stork, P.J. The somatostatin receptors SSTR1 and SSTR2 are coupled to inhibition of adenylyl cyclase in Chinese hamster ovary cells via pertussis toxin-sensitive pathways. Endocrinology 1994, 134, 1277–1285. [Google Scholar] [CrossRef]
  91. Hou, C.; Gilbert, R.L.; Barber, D.L. Subtype-specific signaling mechanisms of somatostatin receptors SSTR1 and SSTR2. J. Biol. Chem. 1994, 269, 10357–10362. [Google Scholar] [CrossRef]
  92. Patel, Y.C.; Srikant, C.B. Subtype selectivity of peptide analogs for all five cloned human somatostatin receptors (hsstr 1–5). Endocrinology 1994, 135, 2814–2817. [Google Scholar] [CrossRef]
  93. Chen, L.; Fitzpatrick, V.D.; Vandlen, R.L.; Tashjian, A.H., Jr. Both overlapping and distinct signaling pathways for somatostatin receptor subtypes SSTR1 and SSTR2 in pituitary cells. J. Biol. Chem. 1997, 272, 18666–18672. [Google Scholar] [CrossRef] [PubMed]
  94. Cervia, D.; Fiorini, S.; Pavan, B.; Biondi, C.; Bagnoli, P. Somatostatin (SRIF) modulates distinct signaling pathways in rat pituitary tumor cells; negative coupling of SRIF receptor subtypes 1 and 2 to arachidonic acid release. Naunyn Schmiedebergs Arch. Pharmacol. 2002, 365, 200–209. [Google Scholar] [CrossRef] [PubMed]
  95. Buchan, A.M.; Lin, C.Y.; Choi, J.; Barber, D.L. Somatostatin, acting at receptor subtype 1, inhibits Rho activity, the assembly of actin stress fibers, and cell migration. J. Biol. Chem. 2002, 277, 28431–28438. [Google Scholar] [CrossRef] [PubMed]
  96. Duluc, C.; Moatassim-Billah, S.; Chalabi-Dchar, M.; Perraud, A.; Samain, R.; Breibach, F.; Gayral, M.; Cordelier, P.; Delisle, M.B.; Bousquet-Dubouch, M.P.; et al. Pharmacological targeting of the protein synthesis mTOR/4E-BP1 pathway in cancer-associated fibroblasts abrogates pancreatic tumour chemoresistance. EMBO Mol. Med. 2015, 7, 735–753. [Google Scholar] [CrossRef]
  97. Florio, T.; Yao, H.; Carey, K.D.; Dillon, T.J.; Stork, P.J. Somatostatin activation of mitogen-activated protein kinase via somatostatin receptor 1 (SSTR1). Mol. Endocrinol. 1999, 13, 24–37. [Google Scholar] [CrossRef] [PubMed]
  98. Reubi, J.C.; Waser, B.; Liu, Q.; Laissue, J.A.; Schonbrunn, A. Subcellular distribution of somatostatin sst2A receptors in human tumors of the nervous and neuroendocrine systems: Membranous versus intracellular location. J. Clin. Endocrinol. Metab. 2000, 85, 3882–3891. [Google Scholar] [CrossRef]
  99. Gugger, M.; Waser, B.; Kappeler, A.; Schonbrunn, A.; Reubi, J.C. Cellular detection of sst2A receptors in human gastrointestinal tissue. Gut 2004, 53, 1431–1436. [Google Scholar] [CrossRef]
  100. Peineau, S.; Guimiot, F.; Csaba, Z.; Jacquier, S.; Fafouri, A.; Schwendimann, L.; de Roux, N.; Schulz, S.; Gressens, P.; Auvin, S.; et al. Somatostatin receptors type 2 and 5 expression and localization during human pituitary development. Endocrinology 2014, 155, 33–39. [Google Scholar] [CrossRef]
  101. Schindler, M.; Holloway, S.; Humphrey, P.P.; Waldvogel, H.; Faull, R.L.; Berger, W.; Emson, P.C. Localization of the somatostatin sst2(a) receptor in human cerebral cortex, hippocampus and cerebellum. Neuroreport 1998, 9, 521–525. [Google Scholar] [CrossRef]
  102. Csaba, Z.; Pirker, S.; Lelouvier, B.; Simon, A.; Videau, C.; Epelbaum, J.; Czech, T.; Baumgartner, C.; Sperk, G.; Dournaud, P. Somatostatin receptor type 2 undergoes plastic changes in the human epileptic dentate gyrus. J. Neuropathol. Exp. Neurol. 2005, 64, 956–969. [Google Scholar] [CrossRef]
  103. Shi, T.J.; Xiang, Q.; Zhang, M.D.; Barde, S.; Kai-Larsen, Y.; Fried, K.; Josephson, A.; Glück, L.; Deyev, S.M.; Zvyagin, A.V.; et al. Somatostatin and its 2A receptor in dorsal root ganglia and dorsal horn of mouse and human: Expression, trafficking and possible role in pain. Mol. Pain. 2014, 10, 12. [Google Scholar] [CrossRef] [PubMed]
  104. Lupp, A.; Hunder, A.; Petrich, A.; Nagel, F.; Doll, C.; Schulz, S. Reassessment of sst(5) somatostatin receptor expression in normal and neoplastic human tissues using the novel rabbit monoclonal antibody UMB-4. Neuroendocrinology 2011, 94, 255–264. [Google Scholar] [CrossRef] [PubMed]
  105. Unger, N.; Ueberberg, B.; Schulz, S.; Saeger, W.; Mann, K.; Petersenn, S. Differential expression of somatostatin receptor subtype 1–5 proteins in numerous human normal tissues. Exp. Clin. Endocrinol. Diabetes 2012, 120, 482–489. [Google Scholar] [CrossRef]
  106. Karasawa, H.; Yakabi, S.; Wang, L.; Stengel, A.; Rivier, J.; Taché, Y. Brain somatostatin receptor 2 mediates the dipsogenic effect of central somatostatin and cortistatin in rats: Role in drinking behavior. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2014, 307, R793–R801. [Google Scholar] [CrossRef] [PubMed]
  107. Law, S.F.; Yasuda, K.; Bell, G.I.; Reisine, T. Gi alpha 3 and G(o) alpha selectively associate with the cloned somatostatin receptor subtype SSTR2. J. Biol. Chem. 1993, 268, 10721–10727. [Google Scholar] [CrossRef] [PubMed]
  108. Kim, J.K.; Kwon, O.; Kim, J.; Kim, E.-K.; Park, H.K.; Lee, J.E.; Kim, K.L.; Choi, J.W.; Lim, S.; Seok, H.; et al. PDZ Domain-containing 1 (PDZK1) Protein Regulates Phospholipase C-β3 (PLC-β3)-specific Activation of Somatostatin by Forming a Ternary Complex with PLC-β3 and Somatostatin Receptors*. J. Biol. Chem. 2012, 287, 21012–21024. [Google Scholar] [CrossRef]
  109. Buscail, L.; Delesque, N.; Estève, J.P.; Saint-Laurent, N.; Prats, H.; Clerc, P.; Robberecht, P.; Bell, G.I.; Liebow, C.; Schally, A.V.; et al. Stimulation of tyrosine phosphatase and inhibition of cell proliferation by somatostatin analogues: Mediation by human somatostatin receptor subtypes SSTR1 and SSTR2. Proc. Natl. Acad. Sci. USA 1994, 91, 2315–2319. [Google Scholar] [CrossRef]
  110. Reardon, D.B.; Dent, P.; Wood, S.L.; Kong, T.; Sturgill, T.W. Activation in vitro of somatostatin receptor subtypes 2, 3, or 4 stimulates protein tyrosine phosphatase activity in membranes from transfected Ras-transformed NIH 3T3 cells: Coexpression with catalytically inactive SHP-2 blocks responsiveness. Mol. Endocrinol. 1997, 11, 1062–1069. [Google Scholar] [CrossRef]
  111. Pagès, P.; Benali, N.; Saint-Laurent, N.; Estève, J.P.; Schally, A.V.; Tkaczuk, J.; Vaysse, N.; Susini, C.; Buscail, L. sst2 somatostatin receptor mediates cell cycle arrest and induction of p27(Kip1). Evidence for the role of SHP-1. J. Biol. Chem. 1999, 274, 15186–15193. [Google Scholar] [CrossRef]
  112. Theodoropoulou, M.; Zhang, J.; Laupheimer, S.; Paez-Pereda, M.; Erneux, C.; Florio, T.; Pagotto, U.; Stalla, G.K. Octreotide, a somatostatin analogue, mediates its antiproliferative action in pituitary tumor cells by altering phosphatidylinositol 3-kinase signaling and inducing Zac1 expression. Cancer Res. 2006, 66, 1576–1582. [Google Scholar] [CrossRef] [PubMed]
  113. Shi, R.; Redman, P.; Ghose, D.; Hwang, H.; Liu, Y.; Ren, X.; Ding, L.J.; Liu, M.; Jones, K.J.; Xu, W. Shank Proteins Differentially Regulate Synaptic Transmission. eNeuro 2017, 4, 1–12. [Google Scholar] [CrossRef] [PubMed]
  114. Pola, S.; Cattaneo, M.G.; Vicentini, L.M. Anti-migratory and anti-invasive effect of somatostatin in human neuroblastoma cells: Involvement of Rac and MAP kinase activity. J. Biol. Chem. 2003, 278, 40601–40606. [Google Scholar] [CrossRef]
  115. Torrisani, J.; Hanoun, N.; Laurell, H.; Lopez, F.; Maoret, J.J.; Souque, A.; Susini, C.; Cordelier, P.; Buscail, L. Identification of an upstream promoter of the human somatostatin receptor, hSSTR2, which is controlled by epigenetic modifications. Endocrinology 2008, 149, 3137–3147. [Google Scholar] [CrossRef] [PubMed]
  116. Shen, Z.; Chen, X.; Li, Q.; Zhou, C.; Li, J.; Ye, H.; Duan, S. SSTR2 promoter hypermethylation is associated with the risk and progression of laryngeal squamous cell carcinoma in males. Diagn. Pathol. 2016, 11, 10. [Google Scholar] [CrossRef]
  117. Lehmann, A.; Kliewer, A.; Schütz, D.; Nagel, F.; Stumm, R.; Schulz, S. Carboxyl-terminal multi-site phosphorylation regulates internalization and desensitization of the human sst2 somatostatin receptor. Mol. Cell Endocrinol. 2014, 387, 44–51. [Google Scholar] [CrossRef]
  118. Lesche, S.; Lehmann, D.; Nagel, F.; Schmid, H.A.; Schulz, S. Differential Effects of Octreotide and Pasireotide on Somatostatin Receptor Internalization and Trafficking in Vitro. J. Clin. Endocrinol. Metab. 2009, 94, 654–661. [Google Scholar] [CrossRef] [PubMed]
  119. Günther, T.; Culler, M.; Schulz, S. Research Resource: Real-Time Analysis of Somatostatin and Dopamine Receptor Signaling in Pituitary Cells Using a Fluorescence-Based Membrane Potential Assay. Mol. Endocrinol. 2016, 30, 479–490. [Google Scholar] [CrossRef] [PubMed]
  120. Sharma, K.; Patel, Y.C.; Srikant, C.B. Subtype-selective induction of wild-type p53 and apoptosis, but not cell cycle arrest, by human somatostatin receptor 3. Mol. Endocrinol. 1996, 10, 1688–1696. [Google Scholar] [CrossRef]
  121. Tulipano, G.; Stumm, R.; Pfeiffer, M.; Kreienkamp, H.J.; Höllt, V.; Schulz, S. Differential beta-arrestin trafficking and endosomal sorting of somatostatin receptor subtypes. J. Biol. Chem. 2004, 279, 21374–21382. [Google Scholar] [CrossRef]
  122. Einstein, E.B.; Patterson, C.A.; Hon, B.J.; Regan, K.A.; Reddi, J.; Melnikoff, D.E.; Mateer, M.J.; Schulz, S.; Johnson, B.N.; Tallent, M.K. Somatostatin signaling in neuronal cilia is critical for object recognition memory. J. Neurosci. 2010, 30, 4306–4314. [Google Scholar] [CrossRef]
  123. Aourz, N.; Portelli, J.; Coppens, J.; De Bundel, D.; Di Giovanni, G.; Van Eeckhaut, A.; Michotte, Y.; Smolders, I. Cortistatin-14 mediates its anticonvulsant effects via sst2 and sst3 but not ghrelin receptors. CNS Neurosci. Ther. 2014, 20, 662–670. [Google Scholar] [CrossRef]
  124. Farrell, S.R.; Raymond, I.D.; Foote, M.; Brecha, N.C.; Barnes, S. Modulation of voltage-gated ion channels in rat retinal ganglion cells mediated by somatostatin receptor subtype 4. J. Neurophysiol. 2010, 104, 1347–1354. [Google Scholar] [CrossRef]
  125. Fehlmann, D.; Langenegger, D.; Schuepbach, E.; Siehler, S.; Feuerbach, D.; Hoyer, D. Distribution and characterisation of somatostatin receptor mRNA and binding sites in the brain and periphery. J. Physiol. 2000, 94, 265–281. [Google Scholar] [CrossRef] [PubMed]
  126. Ludvigsen, E.; Carlsson, C.; Tiensuu Janson, E.; Sandler, S.; Stridsberg, M. Somatostatin receptor 1–5; expression profiles during rat development. Ups. J. Med. Sci. 2015, 120, 157–168. [Google Scholar] [CrossRef] [PubMed]
  127. Panetta, R.; Patel, Y.C. Expression of mRNA for all five human somatostatin receptors (hSSTR1–5) in pituitary tumors. Life Sci. 1995, 56, 333–342. [Google Scholar] [CrossRef] [PubMed]
  128. Reubi, J.C.; Waser, B.; Schaer, J.C.; Laissue, J.A. Somatostatin receptor sst1-sst5 expression in normal and neoplastic human tissues using receptor autoradiography with subtype-selective ligands. Eur. J. Nucl. Med. 2001, 28, 836–846. [Google Scholar] [CrossRef]
  129. Raynor, K.; Lucki, I.; Reisine, T. Somatostatin receptors in the nucleus accumbens selectively mediate the stimulatory effect of somatostatin on locomotor activity in rats. J. Pharmacol. Exp. Ther. 1993, 265, 67–73. [Google Scholar] [PubMed]
  130. Scheich, B.; Gaszner, B.; Kormos, V.; László, K.; Ádori, C.; Borbély, É.; Hajna, Z.; Tékus, V.; Bölcskei, K.; Ábrahám, I.; et al. Somatostatin receptor subtype 4 activation is involved in anxiety and depression-like behavior in mouse models. Neuropharmacology 2016, 101, 204–215. [Google Scholar] [CrossRef]
  131. Scheich, B.; Csekő, K.; Borbély, É.; Ábrahám, I.; Csernus, V.; Gaszner, B.; Helyes, Z. Higher susceptibility of somatostatin 4 receptor gene-deleted mice to chronic stress-induced behavioral and neuroendocrine alterations. Neuroscience 2017, 346, 320–336. [Google Scholar] [CrossRef]
  132. Gorham, L.; Just, S.; Doods, H. Somatostatin 4 receptor activation modulates G-protein coupled inward rectifying potassium channels and voltage stimulated calcium signals in dorsal root ganglion neurons. Eur. J. Pharmacol. 2014, 736, 101–106. [Google Scholar] [CrossRef]
  133. Møller, L.N.; Stidsen, C.E.; Hartmann, B.; Holst, J.J. Somatostatin receptors. Biochim. Biophys. Acta 2003, 1616, 1–84. [Google Scholar] [CrossRef]
  134. Peverelli, E.; Lania, A.G.; Mantovani, G.; Beck-Peccoz, P.; Spada, A. Characterization of intracellular signaling mediated by human somatostatin receptor 5: Role of the DRY motif and the third intracellular loop. Endocrinology 2009, 150, 3169–3176. [Google Scholar] [CrossRef]
  135. Peverelli, E.; Busnelli, M.; Vitali, E.; Giardino, E.; Galés, C.; Lania, A.G.; Beck-Peccoz, P.; Chini, B.; Mantovani, G.; Spada, A. Specific roles of G(i) protein family members revealed by dissecting SST5 coupling in human pituitary cells. J. Cell Sci. 2013, 126, 638–644. [Google Scholar] [CrossRef] [PubMed]
  136. van der Hoek, J.; Lamberts, S.W.; Hofland, L.J. The somatostatin receptor subtype 5 in neuroendocrine tumours. Expert Opin. Investig. Drugs 2010, 19, 385–399. [Google Scholar] [CrossRef] [PubMed]
  137. Wilkinson, G.F.; Feniuk, W.; Humphrey, P.P. Characterization of human recombinant somatostatin sst5 receptors mediating activation of phosphoinositide metabolism. Br. J. Pharmacol. 1997, 121, 91–96. [Google Scholar] [CrossRef] [PubMed]
  138. Wilkinson, G.F.; Feniuk, W.; Humphrey, P.P. Homologous and heterologous desensitisation of somatostatin-induced increases in intracellular Ca2+ and inositol 1,4,5-trisphosphate in CHO-K1 cells expressing human recombinant somatostatin sst5 receptors. Eur. J. Pharmacol. 1997, 340, 277–285. [Google Scholar] [CrossRef] [PubMed]
  139. Tallent, M.; Liapakis, G.; O’Carroll, A.M.; Lolait, S.J.; Dichter, M.; Reisine, T. Somatostatin receptor subtypes SSTR2 and SSTR5 couple negatively to an L-type Ca2+ current in the pituitary cell line AtT-20. Neuroscience 1996, 71, 1073–1081. [Google Scholar] [CrossRef]
  140. Cordelier, P.; Estève, J.-P.; Bousquet, C.; Delesque, N.; O’Carroll, A.-M.; Schally, A.V.; Vaysse, N.; Susini, C.; Buscail, L. Characterization of the antiproliferative signal mediated by the somatostatin receptor subtype sst5. Proc. Natl. Acad. Sci. USA 1997, 94, 9343–9348. [Google Scholar] [CrossRef]
  141. Ballarè, E.; Persani, L.; Lania, A.G.; Filopanti, M.; Giammona, E.; Corbetta, S.; Mantovani, S.; Arosio, M.; Beck-Peccoz, P.; Faglia, G.; et al. Mutation of somatostatin receptor type 5 in an acromegalic patient resistant to somatostatin analog treatment. J. Clin. Endocrinol. Metab. 2001, 86, 3809–3814. [Google Scholar] [CrossRef]
  142. Komatsuzaki, K.; Terashita, K.; Kinane, T.B.; Nishimoto, I. Somatostatin type V receptor activates c-Jun N-terminal kinases via Galpha(12) family G proteins. Biochem. Biophys. Res. Commun. 2001, 289, 1211–1217. [Google Scholar] [CrossRef]
  143. Peverelli, E.; Mantovani, G.; Calebiro, D.; Doni, A.; Bondioni, S.; Lania, A.; Beck-Peccoz, P.; Spada, A. The third intracellular loop of the human somatostatin receptor 5 is crucial for arrestin binding and receptor internalization after somatostatin stimulation. Mol. Endocrinol. 2008, 22, 676–688. [Google Scholar] [CrossRef]
  144. Ben-Shlomo, A.; Pichurin, O.; Barshop, N.J.; Wawrowsky, K.A.; Taylor, J.; Culler, M.D.; Chesnokova, V.; Liu, N.A.; Melmed, S. Selective regulation of somatostatin receptor subtype signaling: Evidence for constitutive receptor activation. Mol. Endocrinol. 2007, 21, 2565–2578. [Google Scholar] [CrossRef]
  145. Ben-Shlomo, A.; Zhou, C.; Pichurin, O.; Chesnokova, V.; Liu, N.A.; Culler, M.D.; Melmed, S. Constitutive somatostatin receptor activity determines tonic pituitary cell response. Mol. Endocrinol. 2009, 23, 337–348. [Google Scholar] [CrossRef] [PubMed]
  146. Riaz, H.; Liang, A.; Khan, M.K.; Dong, P.; Han, L.; Shahzad, M.; Chong, Z.; Ahmad, S.; Hua, G.; Yang, L. Somatostatin and its receptors: Functional regulation in the development of mice Sertoli cells. J. Steroid Biochem. Mol. Biol. 2013, 138, 257–266. [Google Scholar] [CrossRef] [PubMed]
  147. Duran-Prado, M.; Morell, M.; Delgado-Maroto, V.; Castaño, J.P.; Aneiros-Fernandez, J.; de Lecea, L.; Culler, M.D.; Hernandez-Cortes, P.; O’Valle, F.; Delgado, M. Cortistatin inhibits migration and proliferation of human vascular smooth muscle cells and decreases neointimal formation on carotid artery ligation. Circ. Res. 2013, 112, 1444–1455. [Google Scholar] [CrossRef]
  148. Dimitrakopoulou-Strauss, A.; Georgoulias, V.; Eisenhut, M.; Herth, F.; Koukouraki, S.; Mäcke, H.R.; Haberkorn, U.; Strauss, L.G. Quantitative assessment of SSTR2 expression in patients with non-small cell lung cancer using 68Ga-DOTATOC PET and comparison with 18F-FDG PET. Eur. J. Nucl. Med. Mol. Imaging 2006, 33, 823–830. [Google Scholar] [CrossRef] [PubMed]
  149. Muscarella, L.A.; D’Alessandro, V.; la Torre, A.; Copetti, M.; De Cata, A.; Parrella, P.; Sperandeo, M.; Pellegrini, F.; Frusciante, V.; Maiello, E.; et al. Gene expression of somatostatin receptor subtypes SSTR2a, SSTR3 and SSTR5 in peripheral blood of neuroendocrine lung cancer affected patients. Cell. Oncol. 2011, 34, 435–441. [Google Scholar] [CrossRef]
  150. Klomp, M.J.; Refardt, J.; van Koetsveld, P.M.; Campana, C.; Dalm, S.U.; Dogan, F.; van Velthuysen, M.F.; Feelders, R.A.; de Herder, W.W.; Hofland, J.; et al. Epigenetic regulation of SST(2) expression in small intestinal neuroendocrine tumors. Front. Endocrinol. 2023, 14, 1184436. [Google Scholar] [CrossRef]
  151. Chang, T.K.; Su, W.C.; Chen, Y.C.; Chen, P.J.; Li, C.C.; Yeh, Y.S.; Huang, C.W.; Tsai, H.L.; Wang, J.Y. Unknown-primary neuroendocrine neoplasms diagnosed by short-acting somatostatin test: Case series in one institution. Exp. Ther. Med. 2023, 25, 9. [Google Scholar] [CrossRef]
  152. Burns, L.; Naimi, B.; Ronan, M.; Xu, H.; Weber, H.C. Report of a Novel Molecular Profile in Malignant Insulinoma. J. Clin. Med. 2023, 12, 1280. [Google Scholar] [CrossRef] [PubMed]
  153. Waser, B.; Cescato, R.; Liu, Q.; Kao, Y.J.; Körner, M.; Christ, E.; Schonbrunn, A.; Reubi, J.C. Phosphorylation of sst2 receptors in neuroendocrine tumors after octreotide treatment of patients. Am. J. Pathol. 2012, 180, 1942–1949. [Google Scholar] [CrossRef] [PubMed]
  154. Righi, L.; Volante, M.; Tavaglione, V.; Billè, A.; Daniele, L.; Angusti, T.; Inzani, F.; Pelosi, G.; Rindi, G.; Papotti, M. Somatostatin receptor tissue distribution in lung neuroendocrine tumours: A clinicopathologic and immunohistochemical study of 218 ‘clinically aggressive’ cases. Ann. Oncol. 2010, 21, 548–555. [Google Scholar] [CrossRef]
  155. Lapa, C.; Hänscheid, H.; Wild, V.; Pelzer, T.; Schirbel, A.; Werner, R.A.; Droll, S.; Herrmann, K.; Buck, A.K.; Lückerath, K. Somatostatin receptor expression in small cell lung cancer as a prognostic marker and a target for peptide receptor radionuclide therapy. Oncotarget 2016, 7, 20033–20040. [Google Scholar] [CrossRef]
  156. Kim, C.; Liu, S.V.; Subramaniam, D.S.; Torres, T.; Loda, M.; Esposito, G.; Giaccone, G. Phase I study of the (177)Lu-DOTA(0)-Tyr(3)-Octreotate (lutathera) in combination with nivolumab in patients with neuroendocrine tumors of the lung. J. Immunother. Cancer 2020, 8, e000980. [Google Scholar] [CrossRef]
  157. Kim, J.Y.; Kim, J.; Kim, Y.-i.; Yang, D.-H.; Yoo, C.; Park, I.J.; Ryoo, B.-Y.; Ryu, J.-S.; Hong, S.-M. Somatostatin receptor 2 (SSTR2) expression is associated with better clinical outcome and prognosis in rectal neuroendocrine tumors. Sci. Rep. 2024, 14, 4047. [Google Scholar] [CrossRef]
  158. Kaemmerer, D.; Träger, T.; Hoffmeister, M.; Sipos, B.; Hommann, M.; Sänger, J.; Schulz, S.; Lupp, A. Inverse expression of somatostatin and CXCR4 chemokine receptors in gastroenteropancreatic neuroendocrine neoplasms of different malignancy. Oncotarget 2015, 6, 27566–27579. [Google Scholar] [CrossRef]
  159. Qian, Z.R.; Li, T.; Ter-Minassian, M.; Yang, J.; Chan, J.A.; Brais, L.K.; Masugi, Y.; Thiaglingam, A.; Brooks, N.; Nishihara, R. Association between somatostatin receptor expression and clinical outcomes in neuroendocrine tumors. Pancreas 2016, 45, 1386. [Google Scholar] [CrossRef] [PubMed]
  160. Mizutani, G.; Nakanishi, Y.; Watanabe, N.; Honma, T.; Obana, Y.; Seki, T.; Ohni, S.; Nemoto, N. Expression of Somatostatin Receptor (SSTR) Subtypes (SSTR-1, 2A, 3, 4 and 5) in Neuroendocrine Tumors Using Real-time RT-PCR Method and Immunohistochemistry. Acta Histochem. Cytochem. 2012, 45, 167–176. [Google Scholar] [CrossRef]
  161. Schulz, S.; Pauli, S.U.; Schulz, S.; Händel, M.; Dietzmann, K.; Firsching, R.; Höllt, V. Immunohistochemical determination of five somatostatin receptors in meningioma reveals frequent overexpression of somatostatin receptor subtype sst2A. Clin. Cancer Res. 2000, 6, 1865–1874. [Google Scholar]
  162. Agopiantz, M.; Carnot, M.; Denis, C.; Martin, E.; Gauchotte, G. Hormone Receptor Expression in Meningiomas: A Systematic Review. Cancers 2023, 15, 980. [Google Scholar] [CrossRef]
  163. Wu, W.; Zhou, Y.; Wang, Y.; Liu, L.; Lou, J.; Deng, Y.; Zhao, P.; Shao, A. Clinical Significance of Somatostatin Receptor (SSTR) 2 in Meningioma. Front. Oncol. 2020, 10, 1633. [Google Scholar] [CrossRef] [PubMed]
  164. Leijon, H.; Remes, S.; Hagström, J.; Louhimo, J.; Mäenpää, H.; Schalin-Jäntti, C.; Miettinen, M.; Haglund, C.; Arola, J. Variable somatostatin receptor subtype expression in 151 primary pheochromocytomas and paragangliomas. Hum. Pathol. 2019, 86, 66–75. [Google Scholar] [CrossRef] [PubMed]
  165. Shen, Y.; Luo, Y.; Li, M.; Luo, R.; Chen, L.; Gao, X.; Jiang, J.; Liu, Y.; Lu, Z.; Zhang, J. Somatostatin receptor subtype 2A expression and genetics in 184 paragangliomas: A single center retrospective observational study. Endocrine 2024, 85, 398–406. [Google Scholar] [CrossRef]
  166. Parvizi, N.; Alsafi, A.; Vergine, M.; Ramachandran, R.; Martin, N.; Palazzo, F.; Pinato, D.; Sharma, R.; Meeran, K.; Dina, R. Expression of Somatostatin Receptors in Phaeochromocytoma and Paragangliomas. In Proceedings of the Endocrine Abstracts; Bioscientifica: Bristol, UK, 2012. [Google Scholar]
  167. Kaemmerer, D.; Sänger, J.; Arsenic, R.; D’Haese, J.G.; Neumann, J.; Schmitt-Graeff, A.; Wirtz, R.M.; Schulz, S.; Lupp, A. Evaluation of somatostatin, CXCR4 chemokine and endothelin A receptor expression in a large set of paragangliomas. Oncotarget 2017, 8, 89958–89969. [Google Scholar] [CrossRef]
  168. Hertelendi, M.; Belguenani, O.; Cherfi, A.; Folitar, I.; Kollar, G.; Polack, B.D. Efficacy and Safety of [177Lu]Lu-DOTA-TATE in Adults with Inoperable or Metastatic Somatostatin Receptor-Positive Pheochromocytomas/Paragangliomas, Bronchial and Unknown Origin Neuroendocrine Tumors, and Medullary Thyroid Carcinoma: A Systematic Literature Review. Biomedicines 2023, 11, 1024. [Google Scholar] [CrossRef] [PubMed]
  169. Adelman, D.T.; Liebert, K.J.; Nachtigall, L.B.; Lamerson, M.; Bakker, B. Acromegaly: The disease, its impact on patients, and managing the burden of long-term treatment. Int. J. Gen. Med. 2013, 6, 31–38. [Google Scholar] [CrossRef] [PubMed]
  170. Gatto, F.; Biermasz, N.R.; Feelders, R.A.; Kros, J.M.; Dogan, F.; van der Lely, A.J.; Neggers, S.J.; Lamberts, S.W.; Pereira, A.M.; Ferone, D.; et al. Low beta-arrestin expression correlates with the responsiveness to long-term somatostatin analog treatment in acromegaly. Eur. J. Endocrinol. 2016, 174, 651–662. [Google Scholar] [CrossRef]
  171. Ferone, D.; de Herder, W.W.; Pivonello, R.; Kros, J.M.; van Koetsveld, P.M.; de Jong, T.; Minuto, F.; Colao, A.; Lamberts, S.W.; Hofland, L.J. Correlation of in vitro and in vivo somatotropic adenoma responsiveness to somatostatin analogs and dopamine agonists with immunohistochemical evaluation of somatostatin and dopamine receptors and electron microscopy. J. Clin. Endocrinol. Metab. 2008, 93, 1412–1417. [Google Scholar] [CrossRef]
  172. Gatto, F.; Feelders, R.A.; van der Pas, R.; Kros, J.M.; Waaijers, M.; Sprij-Mooij, D.; Neggers, S.J.; van der Lelij, A.J.; Minuto, F.; Lamberts, S.W.; et al. Immunoreactivity score using an anti-sst2A receptor monoclonal antibody strongly predicts the biochemical response to adjuvant treatment with somatostatin analogs in acromegaly. J. Clin. Endocrinol. Metab. 2013, 98, E66–E71. [Google Scholar] [CrossRef]
  173. Wildemberg, L.E.; Neto, L.V.; Costa, D.F.; Nasciuti, L.E.; Takiya, C.M.; Alves, L.M.; Rebora, A.; Minuto, F.; Ferone, D.; Gadelha, M.R. Low somatostatin receptor subtype 2, but not dopamine receptor subtype 2 expression predicts the lack of biochemical response of somatotropinomas to treatment with somatostatin analogs. J. Endocrinol. Investig. 2013, 36, 38–43. [Google Scholar] [CrossRef]
  174. Obari, A.; Sano, T.; Ohyama, K.; Kudo, E.; Qian, Z.R.; Yoneda, A.; Rayhan, N.; Mustafizur Rahman, M.; Yamada, S. Clinicopathological features of growth hormone-producing pituitary adenomas: Difference among various types defined by cytokeratin distribution pattern including a transitional form. Endocr. Pathol. 2008, 19, 82–91. [Google Scholar] [CrossRef]
  175. Mayr, B.; Buslei, R.; Theodoropoulou, M.; Stalla, G.K.; Buchfelder, M.; Schöfl, C. Molecular and functional properties of densely and sparsely granulated GH-producing pituitary adenomas. Eur. J. Endocrinol. 2013, 169, 391–400. [Google Scholar] [CrossRef]
  176. Kato, M.; Inoshita, N.; Sugiyama, T.; Tani, Y.; Shichiri, M.; Sano, T.; Yamada, S.; Hirata, Y. Differential expression of genes related to drug responsiveness between sparsely and densely granulated somatotroph adenomas. Endocr. J. 2012, 59, 221–228. [Google Scholar] [CrossRef] [PubMed]
  177. Brzana, J.; Yedinak, C.G.; Gultekin, S.H.; Delashaw, J.B.; Fleseriu, M. Growth hormone granulation pattern and somatostatin receptor subtype 2A correlate with postoperative somatostatin receptor ligand response in acromegaly: A large single center experience. Pituitary 2013, 16, 490–498. [Google Scholar] [CrossRef] [PubMed]
  178. Montella, L.; Ottaviano, M.; Morra, R.; Pietroluongo, E.; De Placido, P.; Tortora, M.; Sorrentino, C.; Facchini, G.; De Placido, S.; Giuliano, M.; et al. The Never-Ending History of Octreotide in Thymic Tumors: A Vintage or A Contemporary Drug? Cancers 2022, 14, 774. [Google Scholar] [CrossRef] [PubMed]
  179. Ferone, D.; van Hagen, M.P.; Kwekkeboom, D.J.; van Koetsveld, P.M.; Mooy, D.M.; Lichtenauer-Kaligis, E.; Schönbrunn, A.; Colao, A.; Lamberts, S.W.; Hofland, L.J. Somatostatin receptor subtypes in human thymoma and inhibition of cell proliferation by octreotide in vitro. J. Clin. Endocrinol. Metab. 2000, 85, 1719–1726. [Google Scholar] [CrossRef]
  180. Anderson, B.; Nostedt, J.; Girgis, S.; Dixon, T.; Agrawal, V.; Wiebe, E.; Senior, P.A.; Shapiro, A.M. Insulinoma or non-insulinoma pancreatogenous hypoglycemia? A diagnostic dilemma. J. Surg. Case Rep. 2016, 2016, rjw188. [Google Scholar] [CrossRef]
  181. Peltola, E.; Vesterinen, T.; Leijon, H.; Hannula, P.; Huhtala, H.; Mäkinen, M.; Nieminen, L.; Pirinen, E.; Rönty, M.; Söderström, M.; et al. Immunohistochemical somatostatin receptor expression in insulinomas. Apmis 2023, 131, 152–160. [Google Scholar] [CrossRef]
  182. Casar-Borota, O.; Heck, A.; Schulz, S.; Nesland, J.M.; Ramm-Pettersen, J.; Lekva, T.; Alafuzoff, I.; Bollerslev, J. Expression of SSTR2a, but not of SSTRs 1, 3, or 5 in somatotroph adenomas assessed by monoclonal antibodies was reduced by octreotide and correlated with the acute and long-term effects of octreotide. J. Clin. Endocrinol. Metab. 2013, 98, E1730–E1739. [Google Scholar] [CrossRef]
  183. Gabalec, F.; Drastikova, M.; Cesak, T.; Netuka, D.; Masopust, V.; Machac, J.; Marek, J.; Cap, J.; Beranek, M. Dopamine 2 and somatostatin 1–5 receptors coexpression in clinically non-functioning pituitary adenomas. Physiol. Res. 2015, 64, 369–377. [Google Scholar] [CrossRef] [PubMed]
  184. Lee, M.; Lupp, A.; Mendoza, N.; Martin, N.; Beschorner, R.; Honegger, J.; Schlegel, J.; Shively, T.; Pulz, E.; Schulz, S.; et al. SSTR3 is a putative target for the medical treatment of gonadotroph adenomas of the pituitary. Endocr. Relat. Cancer 2015, 22, 111–119. [Google Scholar] [CrossRef]
  185. Bertherat, J.; Tenenbaum, F.; Perlemoine, K.; Videau, C.; Alberini, J.L.; Richard, B.; Dousset, B.; Bertagna, X.; Epelbaum, J. Somatostatin receptors 2 and 5 are the major somatostatin receptors in insulinomas: An in vivo and in vitro study. J. Clin. Endocrinol. Metab. 2003, 88, 5353–5360. [Google Scholar] [CrossRef] [PubMed]
  186. Portela-Gomes, G.M.; Stridsberg, M.; Grimelius, L.; Rorstad, O.; Janson, E.T. Differential expression of the five somatostatin receptor subtypes in human benign and malignant insulinomas—Predominance of receptor subtype 4. Endocr. Pathol. 2007, 18, 79–85. [Google Scholar] [CrossRef] [PubMed]
  187. Papotti, M.; Macri’, L.; Pagani, A.; Aloi, F.; Bussolati, G. Quantitation of somatostatin receptor type 2 in neuroendocrine (Merkel cell) carcinoma of the skin by competitive RT-PCR. Endocr. Pathol. 1999, 10, 37–46. [Google Scholar] [CrossRef]
  188. Gardair, C.; Samimi, M.; Touzé, A.; Coursaget, P.; Lorette, G.; Caille, A.; Wierzbicka, E.; Croué, A.; Avenel-Audran, M.; Aubin, F.; et al. Somatostatin Receptors 2A and 5 Are Expressed in Merkel Cell Carcinoma with No Association with Disease Severity. Neuroendocrinology 2015, 101, 223–235. [Google Scholar] [CrossRef]
  189. Anderson, A.; Qazi, J.; Shantha, E.; Takagishi, S.; Iyer, J.G.; Blom, A.; Behnia, F.S.; Vesselle, H.; Nghiem, P.; Bhatia, S. Therapeutic targeting of somatostatin receptors in patients with metastatic Merkel cell carcinoma: A retrospective case series. J. Clin. Oncol. 2015, 33, e20031. [Google Scholar] [CrossRef]
  190. Fagerstedt, K.W.; Vesterinen, T.; Leijon, H.; Sihto, H.; Böhling, T.; Arola, J. Somatostatin receptor expression in Merkel cell carcinoma: Correlation with clinical data. Acta Oncol. 2023, 62, 1001–1007. [Google Scholar] [CrossRef]
  191. Schartinger, V.H.; Falkeis, C.; Laimer, K.; Sprinzl, G.M.; Riechelmann, H.; Rasse, M.; Virgolini, I.; Dudás, J. Neuroendocrine differentiation in head and neck squamous cell carcinoma. J. Laryngol. Otol. 2012, 126, 1261–1270. [Google Scholar] [CrossRef]
  192. Lum, S.S.; Fletcher, W.S.; O’Dorisio, M.S.; Nance, R.W.; Pommier, R.F.; Caprara, M. Distribution and functional significance of somatostatin receptors in malignant melanoma. World J. Surg. 2001, 25, 407–412. [Google Scholar] [CrossRef]
  193. Caruntu, A.; Scheau, C.; Tampa, M.; Georgescu, S.R.; Caruntu, C.; Tanase, C. Complex Interaction among Immune, Inflammatory, and Carcinogenic Mechanisms in the Head and Neck Squamous Cell Carcinoma. Adv. Exp. Med. Biol. 2021, 1335, 11–35. [Google Scholar] [CrossRef] [PubMed]
  194. Zhao, M.; Lu, T.; Bi, G.; Hu, Z.; Liang, J.; Bian, Y.; Feng, M.; Zhan, C. PLK1 regulating chemoradiotherapy sensitivity of esophageal squamous cell carcinoma through pentose phosphate pathway/ferroptosis. Biomed. Pharmacother. 2023, 168, 115711. [Google Scholar] [CrossRef] [PubMed]
  195. Tampa, M.; Georgescu, S.R.; Mitran, C.I.; Mitran, M.I.; Matei, C.; Scheau, C.; Constantin, C.; Neagu, M. Recent Advances in Signaling Pathways Comprehension as Carcinogenesis Triggers in Basal Cell Carcinoma. J. Clin. Med. 2020, 9, 3010. [Google Scholar] [CrossRef]
  196. Mizumoto, A.; Ohashi, S.; Hirohashi, K.; Amanuma, Y.; Matsuda, T.; Muto, M. Molecular Mechanisms of Acetaldehyde-Mediated Carcinogenesis in Squamous Epithelium. Int. J. Mol. Sci. 2017, 18, 1943. [Google Scholar] [CrossRef]
  197. Loh, K.S.; Waser, B.; Tan, L.K.; Ruan, R.S.; Stauffer, E.; Reubi, J.C. Somatostatin receptors in nasopharyngeal carcinoma. Virchows Arch. 2002, 441, 444–448. [Google Scholar] [CrossRef]
  198. Fan, S.; Zheng, H.; Zhan, Y.; Luo, J.; Zang, H.; Wang, H.; Wang, W.; Xu, Y. Somatostatin receptor2 (SSTR2) expression, prognostic implications, modifications and potential therapeutic strategies associates with head and neck squamous cell carcinomas. Crit. Rev. Oncol. Hematol. 2024, 193, 104223. [Google Scholar] [CrossRef]
  199. Harda, K.; Szabo, Z.; Szabo, E.; Olah, G.; Fodor, K.; Szasz, C.; Mehes, G.; Schally, A.V.; Halmos, G. Somatostatin Receptors as Molecular Targets in Human Uveal Melanoma. Molecules 2018, 23, 1535. [Google Scholar] [CrossRef]
  200. Valsecchi, M.E.; Coronel, M.; Intenzo, C.M.; Kim, S.M.; Witkiewicz, A.K.; Sato, T. Somatostatin receptor scintigraphy in patients with metastatic uveal melanoma. Melanoma Res. 2013, 23, 33–39. [Google Scholar] [CrossRef] [PubMed]
  201. Kumar, U.; Grigorakis, S.I.; Watt, H.L.; Sasi, R.; Snell, L.; Watson, P.; Chaudhari, S. Somatostatin receptors in primary human breast cancer: Quantitative analysis of mRNA for subtypes 1–5 and correlation with receptor protein expression and tumor pathology. Breast Cancer Res. Treat. 2005, 92, 175–186. [Google Scholar] [CrossRef]
  202. Zou, Y.; Tan, H.; Zhao, Y.; Zhou, Y.; Cao, L. Expression and selective activation of somatostatin receptor subtypes induces cell cycle arrest in cancer cells. Oncol. Lett. 2019, 17, 1723–1731. [Google Scholar] [CrossRef]
  203. Hall, G.H.; Turnbull, L.W.; Richmond, I.; Helboe, L.; Atkin, S.L. Localisation of somatostatin and somatostatin receptors in benign and malignant ovarian tumours. Br. J. Cancer 2002, 87, 86–90. [Google Scholar] [CrossRef]
  204. Schulz, S.; Schmitt, J.; Quednow, C.; Roessner, A.; Weise, W. Immunohistochemical detection of somatostatin receptors in human ovarian tumors. Gynecol. Oncol. 2002, 84, 235–240. [Google Scholar] [CrossRef] [PubMed]
  205. Priyadarshini, S.; Allison, D.B.; Chauhan, A. Comprehensive Assessment of Somatostatin Receptors in Various Neoplasms: A Systematic Review. Pharmaceutics 2022, 14, 1394. [Google Scholar] [CrossRef]
  206. Zhao, Y.; Peng, L.; Li, X.; Zhang, Y. Expression of somatostatin and its receptor 1–5 in endometriotic tissues and cells. Exp. Ther. Med. 2018, 16, 3777–3784. [Google Scholar] [CrossRef]
  207. Chapron, C.; Marcellin, L.; Borghese, B.; Santulli, P. Rethinking mechanisms, diagnosis and management of endometriosis. Nat. Rev. Endocrinol. 2019, 15, 666–682. [Google Scholar] [CrossRef] [PubMed]
  208. Bedaiwy, M.A.; Alfaraj, S.; Yong, P.; Casper, R. New developments in the medical treatment of endometriosis. Fertil. Steril. 2017, 107, 555–565. [Google Scholar] [CrossRef] [PubMed]
  209. Halmos, G.; Schally, A.V.; Sun, B.; Davis, R.; Bostwick, D.G.; Plonowski, A. High expression of somatostatin receptors and messenger ribonucleic acid for its receptor subtypes in organ-confined and locally advanced human prostate cancers. J. Clin. Endocrinol. Metab. 2000, 85, 2564–2571. [Google Scholar]
  210. Sinisi, A.A.; Bellastella, A.; Prezioso, D.; Nicchio, M.R.; Lotti, T.; Salvatore, M.; Pasquali, D. Different expression patterns of somatostatin receptor subtypes in cultured epithelial cells from human normal prostate and prostate cancer. J. Clin. Endocrinol. Metab. 1997, 82, 2566–2569. [Google Scholar] [CrossRef]
  211. Kosari, F.; Munz, J.M.; Savci-Heijink, C.D.; Spiro, C.; Klee, E.W.; Kube, D.M.; Tillmans, L.; Slezak, J.; Karnes, R.J.; Cheville, J.C.; et al. Identification of prognostic biomarkers for prostate cancer. Clin. Cancer Res. 2008, 14, 1734–1743. [Google Scholar] [CrossRef]
  212. Pedraza-Arévalo, S.; Hormaechea-Agulla, D.; Gómez-Gómez, E.; Requena, M.J.; Selth, L.A.; Gahete, M.D.; Castaño, J.P.; Luque, R.M. Somatostatin receptor subtype 1 as a potential diagnostic marker and therapeutic target in prostate cancer. Prostate 2017, 77, 1499–1511. [Google Scholar] [CrossRef]
  213. Zhao, J.; Liang, Q.; Cheung, K.F.; Kang, W.; Dong, Y.; Lung, R.W.; Tong, J.H.; To, K.F.; Sung, J.J.; Yu, J. Somatostatin receptor 1, a novel EBV-associated CpG hypermethylated gene, contributes to the pathogenesis of EBV-associated gastric cancer. Br. J. Cancer 2013, 108, 2557–2564. [Google Scholar] [CrossRef] [PubMed]
  214. Werner, C.; Dirsch, O.; Dahmen, U.; Grimm, M.-O.; Schulz, S.; Lupp, A. Evaluation of somatostatin and CXCR4 receptor expression in a large set of prostate cancer samples using tissue microarrays and well-characterized monoclonal antibodies. Transl. Oncol. 2020, 13, 100801. [Google Scholar] [CrossRef] [PubMed]
  215. Hennigs, J.K.; Müller, J.; Adam, M.; Spin, J.M.; Riedel, E.; Graefen, M.; Bokemeyer, C.; Sauter, G.; Huland, H.; Schlomm, T.; et al. Loss of somatostatin receptor subtype 2 in prostate cancer is linked to an aggressive cancer phenotype, high tumor cell proliferation and predicts early metastatic and biochemical relapse. PLoS ONE 2014, 9, e100469. [Google Scholar] [CrossRef]
  216. Romiti, A.; Di Rocco, R.; Milione, M.; Ruco, L.; Ziparo, V.; Zullo, A.; Duranti, E.; Sarcina, I.; Barucca, V.; D’Antonio, C.; et al. Somatostatin receptor subtype 2 A (SSTR2A) and HER2 expression in gastric adenocarcinoma. Anticancer. Res. 2012, 32, 115–119. [Google Scholar]
  217. Hu, C.; Yi, C.; Hao, Z.; Cao, S.; Li, H.; Shao, X.; Zhang, J.; Qiao, T.; Fan, D. The effect of somatostatin and SSTR3 on proliferation and apoptosis of gastric cancer cells. Cancer Biol. Ther. 2004, 3, 726–730. [Google Scholar] [CrossRef]
  218. Chen, W.; Ding, R.; Tang, J.; Li, H.; Chen, C.; Zhang, Y.; Zhang, Q.; Zhu, X. Knocking Out SST Gene of BGC823 Gastric Cancer Cell by CRISPR/Cas9 Enhances Migration, Invasion and Expression of SEMA5A and KLF2. Cancer Manag. Res. 2020, 12, 1313–1321. [Google Scholar] [CrossRef]
  219. Casini Raggi, C.; Calabrò, A.; Renzi, D.; Briganti, V.; Cianchi, F.; Messerini, L.; Valanzano, R.; Cameron Smith, M.; Cortesini, C.; Tonelli, F.; et al. Quantitative evaluation of somatostatin receptor subtype 2 expression in sporadic colorectal tumor and in the corresponding normal mucosa. Clin. Cancer Res. 2002, 8, 419–427. [Google Scholar] [PubMed]
  220. Ferjoux, G.; Bousquet, C.; Cordelier, P.; Benali, N.; Lopez, F.; Rochaix, P.; Buscail, L.; Susini, C. Signal transduction of somatostatin receptors negatively controlling cell proliferation. J. Physiol. 2000, 94, 205–210. [Google Scholar] [CrossRef]
  221. Dasgupta, P. Somatostatin analogues: Multiple roles in cellular proliferation, neoplasia, and angiogenesis. Pharmacol. Ther. 2004, 102, 61–85. [Google Scholar] [CrossRef]
  222. War, S.A.; Kumar, U. Coexpression of human somatostatin receptor-2 (SSTR2) and SSTR3 modulates antiproliferative signaling and apoptosis. J. Mol. Signal. 2012, 7, 5. [Google Scholar] [CrossRef]
  223. Teijeiro, R.; Rios, R.; Costoya, J.A.; Castro, R.; Bello, J.L.; Devesa, J.; Arce, V.M. Activation of human somatostatin receptor 2 promotes apoptosis through a mechanism that is independent from induction of p53. Cell Physiol. Biochem. 2002, 12, 31–38. [Google Scholar] [CrossRef]
  224. Liu, H.L.; Huo, L.; Wang, L. Octreotide inhibits proliferation and induces apoptosis of hepatocellular carcinoma cells. Acta Pharmacol. Sin. 2004, 25, 1380–1386. [Google Scholar] [PubMed]
  225. Lasfer, M.; Vadrot, N.; Schally, A.V.; Nagy, A.; Halmos, G.; Pessayre, D.; Feldmann, G.; Reyl-Desmars, F.J. Potent induction of apoptosis in human hepatoma cell lines by targeted cytotoxic somatostatin analogue AN-238. J. Hepatol. 2005, 42, 230–237. [Google Scholar] [CrossRef] [PubMed]
  226. Chao, T.C.; Chao, H.H.; Chen, M.F.; Lin, J.D. Somatostatin modulates the function of Kupffer cells. Regul. Pept. 1997, 69, 143–149. [Google Scholar] [CrossRef]
  227. Dalm, V.A.; Hofland, L.J.; Lamberts, S.W. Future clinical prospects in somatostatin/cortistatin/somatostatin receptor field. Mol. Cell Endocrinol. 2008, 286, 262–277. [Google Scholar] [CrossRef] [PubMed]
  228. Reynaert, H.; Rombouts, K.; Vandermonde, A.; Urbain, D.; Kumar, U.; Bioulac-Sage, P.; Pinzani, M.; Rosenbaum, J.; Geerts, A. Expression of somatostatin receptors in normal and cirrhotic human liver and in hepatocellular carcinoma. Gut 2004, 53, 1180–1189. [Google Scholar] [CrossRef]
  229. Lang, A.; Sakhnini, E.; Fidder, H.H.; Maor, Y.; Bar-Meir, S.; Chowers, Y. Somatostatin inhibits pro-inflammatory cytokine secretion from rat hepatic stellate cells. Liver Int. 2005, 25, 808–816. [Google Scholar] [CrossRef] [PubMed]
  230. Reubi, J.C.; Waser, B.; Cescato, R.; Gloor, B.; Stettler, C.; Christ, E. Internalized somatostatin receptor subtype 2 in neuroendocrine tumors of octreotide-treated patients. J. Clin. Endocrinol. Metab. 2010, 95, 2343–2350. [Google Scholar] [CrossRef]
  231. Benuck, M.; Marks, N. Differences in the degradation of hypothalamic releasing factors by rat and human serum. Life Sci. 1976, 19, 1271–1276. [Google Scholar] [CrossRef]
  232. Grant, N.; Clark, D.; Garsky, V.; Jaunakais, I.; McGregor, W.; Sarantakis, D. Dissociation of somatostatin effects. Peptides inhibiting the release of growth hormone but not glucagon or insulin in rats. Life Sci. 1976, 19, 629–631. [Google Scholar] [CrossRef]
  233. Brown, M.; Rivier, J.; Vale, W. Somatostatin: Analogs with selected biological activities. Science 1977, 196, 1467–1469. [Google Scholar] [CrossRef]
  234. Meyers, C.; Arimura, A.; Gordin, A.; Fernandez-Durango, R.; Coy, D.H.; Schally, A.V.; Drouin, J.; Ferland, L.; Beaulieu, M.; Labrie, F. Somatostatin analogs which inhibit glucagon and growth hormone more than insulin release. Biochem. Biophys. Res. Commun. 1977, 74, 630–636. [Google Scholar] [CrossRef] [PubMed]
  235. Coy, D.H.; Meyers, C.; Arimura, A.; Schally, A.V.; Redding, T.W. Observations on the growth hormone, insulin, and glucagon release-inhibiting activities of somatostatin analogues. Metabolism 1978, 27, 1407–1410. [Google Scholar] [CrossRef] [PubMed]
  236. Patel, Y.C.; Reichlin, S. Somatostatin in hypothalamus, extrahypothalamic brain, and peripheral tissues of the rat. Endocrinology 1978, 102, 523–530. [Google Scholar] [CrossRef] [PubMed]
  237. Shimon, I.; Yan, X.; Taylor, J.E.; Weiss, M.H.; Culler, M.D.; Melmed, S. Somatostatin receptor (SSTR) subtype-selective analogues differentially suppress in vitro growth hormone and prolactin in human pituitary adenomas. Novel potential therapy for functional pituitary tumors. J. Clin. Investig. 1997, 100, 2386–2392. [Google Scholar] [CrossRef]
  238. Afargan, M.; Janson, E.T.; Gelerman, G.; Rosenfeld, R.; Ziv, O.; Karpov, O.; Wolf, A.; Bracha, M.; Shohat, D.; Liapakis, G.; et al. Novel long-acting somatostatin analog with endocrine selectivity: Potent suppression of growth hormone but not of insulin. Endocrinology 2001, 142, 477–486. [Google Scholar] [CrossRef]
  239. Gomes-Porras, M.; Cárdenas-Salas, J.; Álvarez-Escolá, C. Somatostatin Analogs in Clinical Practice: A Review. Int. J. Mol. Sci. 2020, 21, 1682. [Google Scholar] [CrossRef]
  240. Fattah, S.; Brayden, D.J. Progress in the Formulation and Delivery of Somatostatin Analogs for Acromegaly. Ther. Deliv. 2017, 8, 867–878. [Google Scholar] [CrossRef]
  241. 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]
  242. Bauer, W.; Briner, U.; Doepfner, W.; Haller, R.; Huguenin, R.; Marbach, P.; Petcher, T.J.; Pless, J. SMS 201–995: A very potent and selective octapeptide analogue of somatostatin with prolonged action. Life Sci. 1982, 31, 1133–1140. [Google Scholar] [CrossRef]
  243. Levy, M.J.; Bejon, P.; Barakat, M.; Goadsby, P.J.; Meeran, K. Acromegaly: A Unique Human Headache Model. Headache J. Head. Face Pain. 2003, 43, 794–797. [Google Scholar] [CrossRef] [PubMed]
  244. Bolanowski, M.; Kałużny, M.; Witek, P.; Jawiarczyk-Przybyłowska, A. Pasireotide-a novel somatostatin receptor ligand after 20 years of use. Rev. Endocr. Metab. Disord. 2022, 23, 601–620. [Google Scholar] [CrossRef] [PubMed]
  245. Girard, P.M.; Goldschmidt, E.; Vittecoq, D.; Massip, P.; Gastiaburu, J.; Meyohas, M.C.; Coulaud, J.P.; Schally, A.V. Vapreotide, a somatostatin analogue, in cryptosporidiosis and other AIDS-related diarrhoeal diseases. AIDS 1992, 6, 715–718. [Google Scholar] [CrossRef] [PubMed]
  246. Dimech, J.; Feniuk, W.; Humphrey, P.P. Antagonist effects of seglitide (MK 678) at somatostatin receptors in guinea-pig isolated right atria. Br. J. Pharmacol. 1993, 109, 898–899. [Google Scholar] [CrossRef] [PubMed]
  247. Betoin, F.; Ardid, D.; Herbet, A.; Aumaitre, O.; Kemeny, J.L.; Duchene-Marullaz, P.; Lavarenne, J.; Eschalier, A. Evidence for a central long-lasting antinociceptive effect of vapreotide, an analog of somatostatin, involving an opioidergic mechanism. J. Pharmacol. Exp. Ther. 1994, 269, 7–14. [Google Scholar]
  248. Bétoin, F.; Advenier, C.; Fardin, V.; Wilcox, G.; Lavarenne, J.; Eschalier, A. In vitro and in vivo evidence for a tachykinin NK1 receptor antagonist effect of vapreotide, an analgesic cyclic analog of somatostatin. Eur. J. Pharmacol. 1995, 279, 241–249. [Google Scholar] [CrossRef]
  249. Ritz, M.A.; Drewe, J.; Ziel, A.; Hildebrand, P.; Schneider, P.; Lahlou, N.; Beglinger, C. The effects of vapreotide, a somatostatin analogue, on gastric acidity, gallbladder emptying and hormone release after 1 week of continuous subcutaneous infusion in normal subjects. Br. J. Clin. Pharmacol. 1999, 47, 195–201. [Google Scholar] [CrossRef]
  250. Calès, P.; Masliah, C.; Bernard, B.; Garnier, P.P.; Silvain, C.; Szostak-Talbodec, N.; Bronowicki, J.P.; Ribard, D.; Botta-Fridlund, D.; Hillon, P.; et al. Early administration of vapreotide for variceal bleeding in patients with cirrhosis. N. Engl. J. Med. 2001, 344, 23–28. [Google Scholar] [CrossRef]
  251. Eloubeidi, M.A.; Arguedas, M.R. Early administration of vapreotide for variceal bleeding in patients with cirrhosis. Gastrointest. Endosc. 2002, 55, 135–137. [Google Scholar] [CrossRef]
  252. Veal, N.; Moal, F.; Oberti, F.; Vuillemin, E.; Calés, P. Hemodynamic effects of acute and chronic administration of vapreotide in rats with cirrhosis. Dig. Dis. Sci. 2003, 48, 154–161. [Google Scholar] [CrossRef]
  253. Calès, P. Vapreotide acetate for the treatment of esophageal variceal bleeding. Expert Rev. Gastroenterol. Hepatol. 2008, 2, 185–192. [Google Scholar] [CrossRef] [PubMed]
  254. Fortune, B.E.; Jackson, J.; Leonard, J.; Trotter, J.F. Vapreotide: A somatostatin analog for the treatment of acute variceal bleeding. Expert Opin. Pharmacother. 2009, 10, 2337–2342. [Google Scholar] [CrossRef] [PubMed]
  255. Spitsin, S.; Tuluc, F.; Meshki, J.; Ping Lai, J.; Tustin Iii, R.; Douglas, S.D. Analog of somatostatin vapreotide exhibits biological effects in vitro via interaction with neurokinin-1 receptor. Neuroimmunomodulation 2013, 20, 247–255. [Google Scholar] [CrossRef] [PubMed]
  256. Patch, D.; Burroughs, A. Vapreotide in variceal bleeding. J. Hepatol. 2002, 37, 167–168. [Google Scholar] [CrossRef]
  257. Abdellatif, A.A.H.; Abou-Taleb, H.A.; Abd El Ghany, A.A.; Lutz, I.; Bouazzaoui, A. Targeting of somatostatin receptors expressed in blood cells using quantum dots coated with vapreotide. Saudi Pharm. J. 2018, 26, 1162–1169. [Google Scholar] [CrossRef]
  258. Thakur, M.; Kolan, H.; Rifat, S.; Li, J.; Rux, A.; John, E.; Halmos, G.; Schally, A. Vapreotide labeled with Tc-99m for imaging tumors. Int. J. Oncol. 1996, 9, 445–451. [Google Scholar] [CrossRef]
  259. O’Byrne, K.J.; Dobbs, N.; Propper, D.J.; Braybrooke, J.P.; Koukourakis, M.I.; Mitchell, K.; Woodhull, J.; Talbot, D.C.; Schally, A.V.; Harris, A.L. Phase II study of RC-160 (vapreotide), an octapeptide analogue of somatostatin, in the treatment of metastatic breast cancer. Br. J. Cancer 1999, 79, 1413–1418. [Google Scholar] [CrossRef]
  260. Feng, Q.; Yu, M.Z.; Wang, J.C.; Hou, W.J.; Gao, L.Y.; Ma, X.F.; Pei, X.W.; Niu, Y.J.; Liu, X.Y.; Qiu, C.; et al. Synergistic inhibition of breast cancer by co-delivery of VEGF siRNA and paclitaxel via vapreotide-modified core-shell nanoparticles. Biomaterials 2014, 35, 5028–5038. [Google Scholar] [CrossRef] [PubMed]
  261. Dasgupta, P.; Gűnther, T.; Schulz, S. Pharmacological Characterization of Veldoreotide as a Somatostatin Receptor 4 Agonist. Life 2021, 11, 1075. [Google Scholar] [CrossRef]
  262. Veber, D.F.; Saperstein, R.; Nutt, R.F.; Freidinger, R.M.; Brady, S.F.; Curley, P.; Perlow, D.S.; Paleveda, W.J.; Colton, C.D.; Zacchei, A.G.; et al. A super active cyclic hexapeptide analog of somatostatin. Life Sci. 1984, 34, 1371–1378. [Google Scholar] [CrossRef]
  263. Battershill, P.E.; Clissold, S.P. Octreotide. A review of its pharmacodynamic and pharmacokinetic properties, and therapeutic potential in conditions associated with excessive peptide secretion. Drugs 1989, 38, 658–702. [Google Scholar] [CrossRef] [PubMed]
  264. Novartis Pharmaceuticals Corporation. Sandostatin. In Octreotide Acetate Injection; Novartis Pharmaceuticals Corporation: East Hanover, NJ, USA, 2002. [Google Scholar]
  265. Cendros, J.M.; Peraire, C.; Trocóniz, I.F.; Obach, R. Pharmacokinetics and population pharmacodynamic analysis of lanreotide Autogel. Metabolism 2005, 54, 1276–1281. [Google Scholar] [CrossRef] [PubMed]
  266. U.S. Food and Drug Administration. FDA Approved Products: Mycapssa (Octreotide) Oral Delayed-Release Capsules; Center for Drug Evaluation and Research, Ed.; FDA: Silver Spring, MD, USA, 2020; p. 20.
  267. U.S. Food and Drug Administration. FDA Approved Drug Products: Signifor (Pasireotide Diaspartate) Injection, for Subcutaneous Use; Center for Drug Evaluation and Research, Ed.; FDA: Silver Spring, MD, USA, 2012; p. 27.
  268. U.S. Food and Drug Administration. FDA Approved Drug Products: Somatuline Depot (Lanreotide) for Subcutaneous Injection; Center for Drug Evaluation and Research, Ed.; FDA: Silver Spring, MD, USA, 2019; p. 27.
  269. Corica, G.; Ceraudo, M.; Campana, C.; Nista, F.; Cocchiara, F.; Boschetti, M.; Zona, G.; Criminelli, D.; Ferone, D.; Gatto, F. Octreotide-Resistant Acromegaly: Challenges and Solutions. Ther. Clin. Risk Manag. 2020, 16, 379–391. [Google Scholar] [CrossRef] [PubMed]
  270. Gadelha, M.R.; Wildemberg, L.E.; Kasuki, L. The Future of Somatostatin Receptor Ligands in Acromegaly. J. Clin. Endocrinol. Metab. 2022, 107, 297–308. [Google Scholar] [CrossRef]
  271. Banerjee, I.; Salomon-Estebanez, M.; Shah, P.; Nicholson, J.; Cosgrove, K.E.; Dunne, M.J. Therapies and outcomes of congenital hyperinsulinism-induced hypoglycaemia. Diabet. Med. 2019, 36, 9–21. [Google Scholar] [CrossRef]
  272. Ruuska, T.; Ramírez Escalante, Y.; Vaittinen, S.; Gardberg, M.; Kiviniemi, A.; Marjamäki, P.; Kemppainen, J.; Jyrkkiö, S.; Minn, H. Somatostatin receptor expression in lymphomas: A source of false diagnosis of neuroendocrine tumor at 68Ga-DOTANOC PET/CT imaging. Acta Oncol. 2018, 57, 283–289. [Google Scholar] [CrossRef]
  273. Cives, M.; Strosberg, J.R. Gastroenteropancreatic Neuroendocrine Tumors. CA Cancer J. Clin. 2018, 68, 471–487. [Google Scholar] [CrossRef]
  274. Guida, M.; D’Alò, A.; Mangia, A.; Di Pinto, F.; Sonnessa, M.; Albano, A.; Sciacovelli, A.; Asabella, A.N.; Fucci, L. Somatostatin Receptors in Merkel-Cell Carcinoma: A Therapeutic Opportunity Using Somatostatin Analog alone or in Association with Checkpoint Inhibitors Immunotherapy. A Case Report. Front. Oncol. 2020, 10, 1073. [Google Scholar] [CrossRef]
  275. Akaike, T.; Qazi, J.; Anderson, A.; Behnia, F.S.; Shinohara, M.M.; Akaike, G.; Hippe, D.S.; Thomas, H.; Takagishi, S.R.; Lachance, K.; et al. High somatostatin receptor expression and efficacy of somatostatin analogues in patients with metastatic Merkel cell carcinoma. Br. J. Dermatol. 2021, 184, 319–327. [Google Scholar] [CrossRef]
  276. de Herder, W.W.; Hofland, J. Insulinoma. In Endotext; Feingold, K.R., Anawalt, B., Blackman, M.R., Boyce, A., Chrousos, G., Corpas, E., de Herder, W.W., Dhatariya, K., Dungan, K., Hofland, J., et al., Eds.; MDText.com, Inc.: South Dartmouth, MA, USA, 2000. [Google Scholar]
  277. Caplin, M.E.; Pavel, M.; Ćwikła, J.B.; Phan, A.T.; Raderer, M.; Sedláčková, E.; Cadiot, G.; Wolin, E.M.; Capdevila, J.; Wall, L.; et al. Lanreotide in metastatic enteropancreatic neuroendocrine tumors. N. Engl. J. Med. 2014, 371, 224–233. [Google Scholar] [CrossRef]
  278. Samonakis, D.N.; Moschandreas, J.; Arnaoutis, T.; Skordilis, P.; Leontidis, C.; Vafiades, I.; Kouroumalis, E. Treatment of hepatocellular carcinoma with long acting somatostatin analogues. Oncol. Rep. 2002, 9, 903–907. [Google Scholar] [CrossRef] [PubMed]
  279. Puig-Domingo, M.; Bernabéu, I.; Picó, A.; Biagetti, B.; Gil, J.; Alvarez-Escolá, C.; Jordà, M.; Marques-Pamies, M.; Soldevila, B.; Gálvez, M.-A.; et al. Pasireotide in the Personalized Treatment of Acromegaly. Front. Endocrinol. 2021, 12, 648411. [Google Scholar] [CrossRef] [PubMed]
  280. Silverstein, J.M. Hyperglycemia induced by pasireotide in patients with Cushing’s disease or acromegaly. Pituitary 2016, 19, 536–543. [Google Scholar] [CrossRef]
  281. Kvols, L.K.; Oberg, K.E.; O’Dorisio, T.M.; Mohideen, P.; de Herder, W.W.; Arnold, R.; Hu, K.; Zhang, Y.; Hughes, G.; Anthony, L.; et al. Pasireotide (SOM230) shows efficacy and tolerability in the treatment of patients with advanced neuroendocrine tumors refractory or resistant to octreotide LAR: Results from a phase II study. Endocr. Relat. Cancer 2012, 19, 657–666. [Google Scholar] [CrossRef] [PubMed]
  282. Plöckinger, U.; Wiedenmann, B. Neuroendocrine tumors. Biotherapy. Best. Pract. Res. Clin. Endocrinol. Metab. 2007, 21, 145–162. [Google Scholar] [CrossRef]
  283. Ro, C.; Chai, W.; Yu, V.E.; Yu, R. Pancreatic neuroendocrine tumors: Biology, diagnosis, and treatment. Chin. J. Cancer 2013, 32, 312–324. [Google Scholar] [CrossRef]
  284. Cives, M.; Strosberg, J. The expanding role of somatostatin analogs in gastroenteropancreatic and lung neuroendocrine tumors. Drugs 2015, 75, 847–858. [Google Scholar] [CrossRef]
  285. Hofland, L.J.; Lamberts, S.W.J. Somatostatin receptors and disease: Role of receptor subtypes. Baillière’s Clin. Endocrinol. Metab. 1996, 10, 163–176. [Google Scholar] [CrossRef]
  286. De Conno, F.; Saita, L.; Ripamonti, C.; Ventafridda, V. Subcutaneous octreotide in the treatment of pain in advanced cancer patients. J. Pain Symptom Manag. 1994, 9, 34–38. [Google Scholar] [CrossRef]
  287. Murphy, E.; Prommer, E.E.; Mihalyo, M.; Wilcock, A. Octreotide. J. Pain Symptom Manag. 2010, 40, 142–148. [Google Scholar] [CrossRef]
  288. Lamberts, S.W.J.; Hofland, L.J. ANNIVERSARY REVIEW: Octreotide, 40 years later. Eur. J. Endocrinol. 2019, 181, R173–R183. [Google Scholar] [CrossRef] [PubMed]
  289. Colao, A.; Merola, B.; Ferone, D.; Lombardi, G. Acromegaly. J. Clin. Endocrinol. Metab. 1997, 82, 2777–2781. [Google Scholar] [CrossRef]
  290. Maffei, P.; Tamagno, G.; Nardelli, G.B.; Videau, C.; Menegazzo, C.; Milan, G.; Calcagno, A.; Martini, C.; Vettor, R.; Epelbaum, J.; et al. Effects of octreotide exposure during pregnancy in acromegaly. Clin. Endocrinol. 2010, 72, 668–677. [Google Scholar] [CrossRef]
  291. Borna, R.M.; Jahr, J.S.; Kmiecik, S.; Mancuso, K.F.; Kaye, A.D. Pharmacology of Octreotide: Clinical Implications for Anesthesiologists and Associated Risks. Anesthesiol. Clin. 2017, 35, 327–339. [Google Scholar] [CrossRef]
  292. Boyle, P.J.; Justice, K.; Krentz, A.J.; Nagy, R.J.; Schade, D.S. Octreotide reverses hyperinsulinemia and prevents hypoglycemia induced by sulfonylurea overdoses. J. Clin. Endocrinol. Metab. 1993, 76, 752–756. [Google Scholar] [CrossRef] [PubMed]
  293. Carr, R.; Zed, P.J. Octreotide for sulfonylurea-induced hypoglycemia following overdose. Ann. Pharmacother. 2002, 36, 1727–1732. [Google Scholar] [CrossRef]
  294. Hassan, Z.; Wright, J. Use of octreotide acetate to prevent rebound hypoglycaemia in sulfonylurea overdose. Emerg. Med. J. 2007, 24, 580–581. [Google Scholar] [CrossRef] [PubMed]
  295. Glatstein, M.; Scolnik, D.; Bentur, Y. Octreotide for the treatment of sulfonylurea poisoning. Clin. Toxicol. 2012, 50, 795–804. [Google Scholar] [CrossRef] [PubMed]
  296. Genuth, S. Should sulfonylureas remain an acceptable first-line add-on to metformin therapy in patients with type 2 diabetes? No, it’s time to move on! Diabetes Care 2015, 38, 170–175. [Google Scholar] [CrossRef]
  297. Sola, D.; Rossi, L.; Schianca, G.P.; Maffioli, P.; Bigliocca, M.; Mella, R.; Corlianò, F.; Fra, G.P.; Bartoli, E.; Derosa, G. Sulfonylureas and their use in clinical practice. Arch. Med. Sci. 2015, 11, 840–848. [Google Scholar] [CrossRef]
  298. Vaughan, E.M.; Rueda, J.J.; Samson, S.L.; Hyman, D.J. Reducing the Burden of Diabetes Treatment: A Review of Low-cost Oral Hypoglycemic Medications. Curr. Diabetes Rev. 2020, 16, 851–858. [Google Scholar] [CrossRef] [PubMed]
  299. Khan, M.A.B.; Hashim, M.J.; King, J.K.; Govender, R.D.; Mustafa, H.; Al Kaabi, J. Epidemiology of Type 2 Diabetes—Global Burden of Disease and Forecasted Trends. J. Epidemiol. Glob. Health 2020, 10, 107–111. [Google Scholar] [CrossRef] [PubMed]
  300. Neagu, M.; Constantin, C.; Surcel, M.; Munteanu, A.; Scheau, C.; Savulescu-Fiedler, I.; Caruntu, C. Diabetic neuropathy: A NRF2 disease? J. Diabetes 2023, 1–17. [Google Scholar] [CrossRef] [PubMed]
  301. Groth, C.M.; Banzon, E.R. Octreotide for the treatment of hypoglycemia after insulin glargine overdose. J. Emerg. Med. 2013, 45, 194–198. [Google Scholar] [CrossRef]
  302. Harvey, M.; Cave, G. Octreotide may attenuate absorption and ameliorate toxicity following enteric drug overdose. Med. Hypotheses 2013, 81, 424–425. [Google Scholar] [CrossRef]
  303. James, R.A.; Rhodes, M.; Rose, P.; Kendall-Taylor, P. Biliary colic on abrupt withdrawal of octreotide. Lancet 1991, 338, 1527. [Google Scholar] [CrossRef]
  304. Sadoul, J.L.; Benchimol, D.; Thyss, A.; Freychet, P. Acute pancreatitis following octreotide withdrawal. Am. J. Med. 1991, 90, 763–764. [Google Scholar] [CrossRef]
  305. Sadoul, J.L.; Benchimol, D.; Thyss, A.; Freychet, P. Side-effects of octreotide withdrawal. Lancet 1992, 339, 376. [Google Scholar] [CrossRef]
  306. Trobonjaca, Z.; Radosević-Stasić, B.; Crncević, Z.; Rukavina, D. Modulatory effects of octreotide on anti-CD3 and dexamethasone-induced apoptosis of murine thymocytes. Int. Immunopharmacol. 2001, 1, 1753–1764. [Google Scholar] [CrossRef]
  307. Petrović-Dergović, D.M.; Rakin, A.K.; Dimitrijević, L.A.; Ristovski, J.S.; Kustrimović, N.Z.; Mićić, M.V. Changes in thymus size, cellularity and relation between thymocyte subpopulations in young adult rats induced by somatostatin-14. Neuropeptides 2007, 41, 485–493. [Google Scholar] [CrossRef]
  308. Sedqi, M.; Roy, S.; Mohanraj, D.; Ramakrishnan, S.; Loh, H.H. Activation of rat thymocytes selectively upregulates the expression of somatostatin receptor subtype-1. Biochem. Mol. Biol. Int. 1996, 38, 103–112. [Google Scholar] [PubMed]
  309. Solomou, K.; Ritter, M.A.; Palmer, D.B. Somatostatin is expressed in the murine thymus and enhances thymocyte development. Eur. J. Immunol. 2002, 32, 1550–1559. [Google Scholar] [CrossRef]
  310. Palmieri, G.; Lastoria, S.; Colao, A.; Vergara, E.; Varrella, P.; Biondi, E.; Selleri, C.; Catalano, L.; Lombardi, G.; Bianco, A.R.; et al. Successful treatment of a patient with a thymoma and pure red-cell aplasia with octreotide and prednisone. N. Engl. J. Med. 1997, 336, 263–265. [Google Scholar] [CrossRef] [PubMed]
  311. Zaucha, R.; Zaucha, J.M.; Jassem, J. Resolution of thymoma-related pure red cell aplasia after octreotide treatment. Acta Oncol. 2007, 46, 864–865. [Google Scholar] [CrossRef]
  312. Kertesz, G.P.; Hauser, P.; Varga, P.; Dabasi, G.; Schuler, D.; Garami, M. Advanced pediatric inoperable thymus carcinoma (type C thymoma): Case report on a novel therapeutic approach. J. Pediatr. Hematol. Oncol. 2007, 29, 774–775. [Google Scholar] [CrossRef] [PubMed]
  313. Ito, J.; Sekiya, M.; Miura, K.; Yoshimi, K.; Suzuki, T.; Seyama, K.; Izumi, H.; Uekusa, T.; Takahashi, K. Refractory recurrent thymoma successfully treated with long-acting somatostatin analogue and prednisolone. Intern. Med. 2009, 48, 1061–1064. [Google Scholar] [CrossRef] [PubMed]
  314. Pettit, L.; El-Modir, A. The role of somatostatin analogues in the treatment of advanced malignant thymomas: Case report and review of the literature. Br. J. Radiol. 2011, 84, e7–e10. [Google Scholar] [CrossRef]
  315. Mei, Z.; Wang, H.; Ren, S.; Wei, J.; Gu, Y. Metastatic thymic carcinoid responds to chemoradiation and octreotide: A case report. Medicine 2018, 97, e13286. [Google Scholar] [CrossRef]
  316. Sorejs, O.; Pesek, M.; Finek, J.; Fiala, O. Octreotide in the treatment of malignant thymoma—Case report. Rep. Pract. Oncol. Radiother. 2020, 25, 882–885. [Google Scholar] [CrossRef] [PubMed]
  317. Loehrer, P.J.; Wang, W.; Johnson, D.H.; Aisner, S.C.; Ettinger, D.S. Octreotide alone or with prednisone in patients with advanced thymoma and thymic carcinoma: An Eastern Cooperative Oncology Group Phase II Trial. J. Clin. Oncol. 2004, 22, 293–299. [Google Scholar] [CrossRef]
  318. Kirzinger, L.; Boy, S.; Marienhagen, J.; Schuierer, G.; Neu, R.; Ried, M.; Hofmann, H.S.; Wiebe, K.; Ströbel, P.; May, C.; et al. Octreotide LAR and Prednisone as Neoadjuvant Treatment in Patients with Primary or Locally Recurrent Unresectable Thymic Tumors: A Phase II Study. PLoS ONE 2016, 11, e0168215. [Google Scholar] [CrossRef]
  319. Ottaviano, M.; Damiano, V.; Tortora, M.; Capuano, M.; Perrone, P.; Forino, C.; Matano, E.; Palmieri, G. P1.17-015 Long Acting Octreotide plus Prednisone in Advanced Thymic Epithelial Tumors: A Real Life Clinical Experience. J. Thorac. Oncol. 2017, 12, S2066. [Google Scholar] [CrossRef]
  320. Vortmeyer, A.O.; Huang, S.; Lubensky, I.; Zhuang, Z. Non-islet origin of pancreatic islet cell tumors. J. Clin. Endocrinol. Metab. 2004, 89, 1934–1938. [Google Scholar] [CrossRef] [PubMed]
  321. Cives, M.; Strosberg, J. An update on gastroenteropancreatic neuroendocrine tumors. Oncology 2014, 28, 749–756. [Google Scholar] [PubMed]
  322. Williams, E.D.; Sandler, M. The classification of carcinoid tum ours. Lancet 1963, 1, 238–239. [Google Scholar] [CrossRef] [PubMed]
  323. Papotti, M.; Bongiovanni, M.; Volante, M.; Allìa, E.; Landolfi, S.; Helboe, L.; Schindler, M.; Cole, S.L.; Bussolati, G. Expression of somatostatin receptor types 1–5 in 81 cases of gastrointestinal and pancreatic endocrine tumors. A correlative immunohistochemical and reverse-transcriptase polymerase chain reaction analysis. Virchows Arch. 2002, 440, 461–475. [Google Scholar] [CrossRef]
  324. Faggiano, A. Long-acting somatostatin analogs and well differentiated neuroendocrine tumors: A 20-year-old story. J. Endocrinol. Investig. 2024, 47, 35–46. [Google Scholar] [CrossRef]
  325. Curt, A.M.; Rada Popa Ilie, I.; Căinap, C.; Bălăcescu, O.; Ghervan, C. MicroRNAs and Treatment with Somatostatin Analogs in Gastro-Entero-Pancreatic Neuroendocrine Neoplasms: Challenges in Personalized Medicine. J. Gastrointest. Liver Dis. 2020, 29, 647–659. [Google Scholar] [CrossRef]
  326. Ruggeri, R.M.; Aini, I.; Gay, S.; Grossrubatscher, E.M.; Mancini, C.; Tarsitano, M.G.; Zamponi, V.; Isidori, A.M.; Colao, A.; Faggiano, A. Efficacy and tolerability of somatostatin analogues according to gender in patients with neuroendocrine tumors. Rev. Endocr. Metab. Disord. 2024, 25, 383–398. [Google Scholar] [CrossRef]
  327. Rubin, J.; Ajani, J.; Schirmer, W.; Venook, A.P.; Bukowski, R.; Pommier, R.; Saltz, L.; Dandona, P.; Anthony, L. Octreotide acetate long-acting formulation versus open-label subcutaneous octreotide acetate in malignant carcinoid syndrome. J. Clin. Oncol. 1999, 17, 600–606. [Google Scholar] [CrossRef]
  328. O’Toole, D.; Ducreux, M.; Bommelaer, G.; Wemeau, J.L.; Bouché, O.; Catus, F.; Blumberg, J.; Ruszniewski, P. Treatment of carcinoid syndrome: A prospective crossover evaluation of lanreotide versus octreotide in terms of efficacy, patient acceptability, and tolerance. Cancer 2000, 88, 770–776. [Google Scholar] [CrossRef]
  329. Gade, A.K.; Olariu, E.; Douthit, N.T. Carcinoid Syndrome: A Review. Cureus 2020, 12, e7186. [Google Scholar] [CrossRef] [PubMed]
  330. Rubin de Celis Ferrari, A.C.; Glasberg, J.; Riechelmann, R.P. Carcinoid syndrome: Update on the pathophysiology and treatment. Clinics 2018, 73, e490s. [Google Scholar] [CrossRef]
  331. Kaltsas, G.; Caplin, M.; Davies, P.; Ferone, D.; Garcia-Carbonero, R.; Grozinsky-Glasberg, S.; Hörsch, D.; Tiensuu Janson, E.; Kianmanesh, R.; Kos-Kudla, B.; et al. ENETS Consensus Guidelines for the Standards of Care in Neuroendocrine Tumors: Pre- and Perioperative Therapy in Patients with Neuroendocrine Tumors. Neuroendocrinology 2017, 105, 245–254. [Google Scholar] [CrossRef]
  332. Massimino, K.; Harrskog, O.; Pommier, S.; Pommier, R. Octreotide LAR and bolus octreotide are insufficient for preventing intraoperative complications in carcinoid patients. J. Surg. Oncol. 2013, 107, 842–846. [Google Scholar] [CrossRef]
  333. Condron, M.E.; Pommier, S.J.; Pommier, R.F. Continuous infusion of octreotide combined with perioperative octreotide bolus does not prevent intraoperative carcinoid crisis. Surgery 2016, 159, 358–365. [Google Scholar] [CrossRef]
  334. Strosberg, J.R.; Al-Toubah, T.; El-Haddad, G.; Reidy Lagunes, D.; Bodei, L. Sequencing of Somatostatin-Receptor-Based Therapies in Neuroendocrine Tumor Patients. J. Nucl. Med. 2024, 65, 340–348. [Google Scholar] [CrossRef]
  335. Epelboym, I.; Mazeh, H. Zollinger-Ellison syndrome: Classical considerations and current controversies. Oncologist 2014, 19, 44–50. [Google Scholar] [CrossRef] [PubMed]
  336. Thakker, R.V.; Newey, P.J.; Walls, G.V.; Bilezikian, J.; Dralle, H.; Ebeling, P.R.; Melmed, S.; Sakurai, A.; Tonelli, F.; Brandi, M.L. Clinical practice guidelines for multiple endocrine neoplasia type 1 (MEN1). J. Clin. Endocrinol. Metab. 2012, 97, 2990–3011. [Google Scholar] [CrossRef]
  337. Papotti, M.; Croce, S.; Bellò, M.; Bongiovanni, M.; Allìa, E.; Schindler, M.; Bussolati, G. Expression of somatostatin receptor types 2, 3 and 5 in biopsies and surgical specimens of human lung tumours. Correlation with preoperative octreotide scintigraphy. Virchows Arch. 2001, 439, 787–797. [Google Scholar] [CrossRef]
  338. Lamberts, S.W. The role of somatostatin and its analogs in the diagnosis and treatment of tumors. Endocr. Rev. 1991, 12, 450–478. [Google Scholar] [CrossRef] [PubMed]
  339. Janson, E.T.; Oberg, K. Long-term management of the carcinoid syndrome. Treatment with octreotide alone and in combination with alpha-interferon. Acta Oncol. 1993, 32, 225–229. [Google Scholar] [CrossRef]
  340. Krenning, E.P.; Valkema, R.; Kwekkeboom, D.J.; de Herder, W.W.; van Eijck, C.H.; de Jong, M.; Pauwels, S.; Reubi, J.C. Molecular imaging as in vivo molecular pathology for gastroenteropancreatic neuroendocrine tumors: Implications for follow-up after therapy. J. Nucl. Med. 2005, 46 (Suppl. S1), 76s–82s. [Google Scholar]
  341. Reubi, J.C.; Schaer, J.C.; Waser, B.; Mengod, G. Expression and localization of somatostatin receptor SSTR1, SSTR2, and SSTR3 messenger RNAs in primary human tumors using in situ hybridization. Cancer Res. 1994, 54, 3455–3459. [Google Scholar] [PubMed]
  342. Reubi, J.C.; Kappeler, A.; Waser, B.; Schonbrunn, A.; Laissue, J. Immunohistochemical localization of somatostatin receptor sst2A in human pancreatic islets. J. Clin. Endocrinol. Metab. 1998, 83, 3746–3749. [Google Scholar] [CrossRef]
  343. Janson, E.T.; Stridsberg, M.; Gobl, A.; Westlin, J.E.; Oberg, K. Determination of somatostatin receptor subtype 2 in carcinoid tumors by immunohistochemical investigation with somatostatin receptor subtype 2 antibodies. Cancer Res. 1998, 58, 2375–2378. [Google Scholar] [PubMed]
  344. Kimura, N.; Pilichowska, M.; Date, F.; Kimura, I.; Schindler, M. Immunohistochemical expression of somatostatin type 2A receptor in neuroendocrine tumors. Clin. Cancer Res. 1999, 5, 3483–3487. [Google Scholar]
  345. Reubi, J.C.; Laissue, J.A.; Waser, B.; Steffen, D.L.; Hipkin, R.W.; Schonbrunn, A. Immunohistochemical detection of somatostatin sst2a receptors in the lymphatic, smooth muscular, and peripheral nervous systems of the human gastrointestinal tract: Facts and artifacts. J. Clin. Endocrinol. Metab. 1999, 84, 2942–2950. [Google Scholar] [CrossRef]
  346. Papotti, M.; Croce, S.; Macrì, L.; Funaro, A.; Pecchioni, C.; Schindler, M.; Bussolati, G. Correlative immunohistochemical and reverse transcriptase polymerase chain reaction analysis of somatostatin receptor type 2 in neuroendocrine tumors of the lung. Diagn. Mol. Pathol. 2000, 9, 47–57. [Google Scholar] [CrossRef]
  347. Körner, M.; Eltschinger, V.; Waser, B.; Schonbrunn, A.; Reubi, J.C. Value of immunohistochemistry for somatostatin receptor subtype sst2A in cancer tissues: Lessons from the comparison of anti-sst2A antibodies with somatostatin receptor autoradiography. Am. J. Surg. Pathol. 2005, 29, 1642–1651. [Google Scholar] [CrossRef]
  348. Volante, M.; Brizzi, M.P.; Faggiano, A.; La Rosa, S.; Rapa, I.; Ferrero, A.; Mansueto, G.; Righi, L.; Garancini, S.; Capella, C.; et al. Somatostatin receptor type 2A immunohistochemistry in neuroendocrine tumors: A proposal of scoring system correlated with somatostatin receptor scintigraphy. Mod. Pathol. 2007, 20, 1172–1182. [Google Scholar] [CrossRef] [PubMed]
  349. Tsuta, K.; Wistuba, I.I.; Moran, C.A. Differential expression of somatostatin receptors 1–5 in neuroendocrine carcinoma of the lung. Pathol. Res. Pract. 2012, 208, 470–474. [Google Scholar] [CrossRef]
  350. Raso, M.G.; Bota-Rabassedas, N.; Wistuba, I.I. Pathology and Classification of SCLC. Cancers 2021, 13, 820. [Google Scholar] [CrossRef]
  351. Şen, F.; Sheikh, G.T.; von Hinten, J.; Schindele, A.; Kircher, M.; Dierks, A.; Pfob, C.H.; Serfling, S.E.; Buck, A.K.; Pelzer, T.; et al. In-Vivo Somatostatin-Receptor Expression in Small Cell Lung Cancer as a Prognostic Image Biomarker and Therapeutic Target. Cancers 2023, 15, 3595. [Google Scholar] [CrossRef]
  352. Kruglyak, K.M.; Lin, E.; Ong, F.S. Next-Generation Sequencing and Applications to the Diagnosis and Treatment of Lung Cancer. Adv. Exp. Med. Biol. 2016, 890, 123–136. [Google Scholar] [CrossRef] [PubMed]
  353. Cainap, C.; Balacescu, O.; Cainap, S.S.; Pop, L.A. Next Generation Sequencing Technology in Lung Cancer Diagnosis. Biology 2021, 10, 864. [Google Scholar] [CrossRef]
  354. Shabani Azim, F.; Houri, H.; Ghalavand, Z.; Nikmanesh, B. Next Generation Sequencing in Clinical Oncology: Applications, Challenges and Promises: A Review Article. Iran. J. Public Health 2018, 47, 1453–1457. [Google Scholar] [PubMed]
  355. Perry, R.R.; Vinik, A.I. Endocrine tumors of the gastrointestinal tract. Annu. Rev. Med. 1996, 47, 57–68. [Google Scholar] [CrossRef]
  356. van Essen, M.; Krenning, E.P.; Kam, B.L.; de Herder, W.W.; Feelders, R.A.; Kwekkeboom, D.J. Salvage therapy with (177)Lu-octreotate in patients with bronchial and gastroenteropancreatic neuroendocrine tumors. J. Nucl. Med. 2010, 51, 383–390. [Google Scholar] [CrossRef]
  357. Kwekkeboom, D.J.; Krenning, E.P. Peptide Receptor Radionuclide Therapy in the Treatment of Neuroendocrine Tumors. Hematol. Oncol. Clin. N. Am. 2016, 30, 179–191. [Google Scholar] [CrossRef]
  358. Cives, M.; Strosberg, J. Radionuclide Therapy for Neuroendocrine Tumors. Curr. Oncol. Rep. 2017, 19, 9. [Google Scholar] [CrossRef] [PubMed]
  359. Brabander, T.; van der Zwan, W.A.; Teunissen, J.J.M.; Kam, B.L.R.; Feelders, R.A.; de Herder, W.W.; van Eijck, C.H.J.; Franssen, G.J.H.; Krenning, E.P.; Kwekkeboom, D.J. Long-Term Efficacy, Survival, and Safety of [(177)Lu-DOTA(0),Tyr(3)]octreotate in Patients with Gastroenteropancreatic and Bronchial Neuroendocrine Tumors. Clin. Cancer Res. 2017, 23, 4617–4624. [Google Scholar] [CrossRef] [PubMed]
  360. van der Zwan, W.A.; Brabander, T.; Kam, B.L.R.; Teunissen, J.J.M.; Feelders, R.A.; Hofland, J.; Krenning, E.P.; de Herder, W.W. Salvage peptide receptor radionuclide therapy with [(177)Lu-DOTA,Tyr(3)]octreotate in patients with bronchial and gastroenteropancreatic neuroendocrine tumours. Eur. J. Nucl. Med. Mol. Imaging 2019, 46, 704–717. [Google Scholar] [CrossRef] [PubMed]
  361. Mihalache, O.; Doran, H.; Poiană, C.; Bîrligea, A.; Cîrstea, M.O.; Pătraşcu, T. Pancreatic Neuroendocrine Tumors—Case Series and Literature Review. Chirurgia 2019, 114, 630–638. [Google Scholar] [CrossRef] [PubMed]
  362. Krausz, Y.; Bar-Ziv, J.; de Jong, R.B.; Ish-Shalom, S.; Chisin, R.; Shibley, N.; Glaser, B. Somatostatin-receptor scintigraphy in the management of gastroenteropancreatic tumors. Am. J. Gastroenterol. 1998, 93, 66–70. [Google Scholar] [CrossRef]
  363. Shi, W.; Johnston, C.F.; Buchanan, K.D.; Ferguson, W.R.; Laird, J.D.; Crothers, J.G.; McIlrath, E.M. Localization of neuroendocrine tumours with [111In] DTPA-octreotide scintigraphy (Octreoscan): A comparative study with CT and MR imaging. Qjm 1998, 91, 295–301. [Google Scholar] [CrossRef]
  364. Krausz, Y.; Freedman, N.; Rubinstein, R.; Lavie, E.; Orevi, M.; Tshori, S.; Salmon, A.; Glaser, B.; Chisin, R.; Mishani, E.; et al. 68Ga-DOTA-NOC PET/CT imaging of neuroendocrine tumors: Comparison with 111In-DTPA-octreotide (OctreoScan®). Mol. Imaging Biol. 2011, 13, 583–593. [Google Scholar] [CrossRef]
  365. Hofman, M.S.; Kong, G.; Neels, O.C.; Eu, P.; Hong, E.; Hicks, R.J. High management impact of Ga-68 DOTATATE (GaTate) PET/CT for imaging neuroendocrine and other somatostatin expressing tumours. J. Med. Imaging Radiat. Oncol. 2012, 56, 40–47. [Google Scholar] [CrossRef]
  366. Wild, D.; Bomanji, J.B.; Benkert, P.; Maecke, H.; Ell, P.J.; Reubi, J.C.; Caplin, M.E. Comparison of 68Ga-DOTANOC and 68Ga-DOTATATE PET/CT within patients with gastroenteropancreatic neuroendocrine tumors. J. Nucl. Med. 2013, 54, 364–372. [Google Scholar] [CrossRef] [PubMed]
  367. Sharma, P.; Arora, S.; Mukherjee, A.; Pal, S.; Sahni, P.; Garg, P.; Khadgawat, R.; Thulkar, S.; Bal, C.; Kumar, R. Predictive value of 68Ga-DOTANOC PET/CT in patients with suspicion of neuroendocrine tumors: Is its routine use justified? Clin. Nucl. Med. 2014, 39, 37–43. [Google Scholar] [CrossRef]
  368. Sharma, P.; Arora, S.; Dhull, V.S.; Naswa, N.; Kumar, R.; Ammini, A.C.; Bal, C. Evaluation of 68Ga-DOTANOC PET/CT imaging in a large exclusive population of pancreatic neuroendocrine tumors. Abdom. Imaging 2015, 40, 299–309. [Google Scholar] [CrossRef] [PubMed]
  369. Sharma, P.; Arora, S.; Karunanithi, S.; Khadgawat, R.; Durgapal, P.; Sharma, R.; Kandasamy, D.; Bal, C.; Kumar, R. Somatostatin receptor based PET/CT imaging with 68Ga-DOTA-Nal3-octreotide for localization of clinically and biochemically suspected insulinoma. Q. J. Nucl. Med. Mol. Imaging 2016, 60, 69–76. [Google Scholar]
  370. Eschbach, R.S.; Hofmann, M.; Späth, L.; Sheikh, G.T.; Delker, A.; Lindner, S.; Jurkschat, K.; Wängler, C.; Wängler, B.; Schirrmacher, R.; et al. Comparison of somatostatin receptor expression in patients with neuroendocrine tumours with and without somatostatin analogue treatment imaged with [18F]SiTATE. Front. Oncol. 2023, 13, 992316. [Google Scholar] [CrossRef]
  371. Chahid, Y.; Hashimi, K.; van de Garde, E.M.W.; Klümpen, H.J.; Hendrikse, N.H.; Booij, J.; Verberne, H.J. The Influence of Long-Acting Somatostatin Analogs on 68Ga-DOTATATE Uptake in Patients with Neuroendocrine Tumors. Clin. Nucl. Med. 2023, 48, 757–762. [Google Scholar] [CrossRef]
  372. Poletto, G.; Cecchin, D.; Sperti, S.; Filippi, L.; Realdon, N.; Evangelista, L. Head-to-Head Comparison between Peptide-Based Radiopharmaceutical for PET and SPECT in the Evaluation of Neuroendocrine Tumors: A Systematic Review. Curr. Issues Mol. Biol. 2022, 44, 5516–5530. [Google Scholar] [CrossRef] [PubMed]
  373. Matasar, M.J.; Zelenetz, A.D. Overview of lymphoma diagnosis and management. Radiol. Clin. N. Am. 2008, 46, 175–198. [Google Scholar] [CrossRef] [PubMed]
  374. Zamfir, M.-A.; Moraru, L.; Dobrea, C.; Scheau, A.-E.; Iacob, S.; Moldovan, C.; Scheau, C.; Caruntu, C.; Caruntu, A. Hematologic Malignancies Diagnosed in the Context of the mRNA COVID-19 Vaccination Campaign: A Report of Two Cases. Medicina 2022, 58, 874. [Google Scholar] [CrossRef]
  375. Dalm, V.A.; Hofland, L.J.; Mooy, C.M.; Waaijers, M.A.; van Koetsveld, P.M.; Langerak, A.W.; Staal, F.T.; van der Lely, A.J.; Lamberts, S.W.; van Hagen, M.P. Somatostatin receptors in malignant lymphomas: Targets for radiotherapy? J. Nucl. Med. 2004, 45, 8–16. [Google Scholar]
  376. Ferone, D.; Semino, C.; Boschetti, M.; Cascini, G.L.; Minuto, F.; Lastoria, S. Initial staging of lymphoma with octreotide and other receptor imaging agents. Semin. Nucl. Med. 2005, 35, 176–185. [Google Scholar] [CrossRef]
  377. Raderer, M.; Traub, T.; Formanek, M.; Virgolini, I.; Österreicher, C.; Fiebiger, W.; Penz, M.; Jäger, U.; Pont, J.; Chott, A. Somatostatin-receptor scintigraphy for staging and follow-up of patients with extraintestinal marginal zone B-cell lymphoma of the mucosa associated lymphoid tissue (MALT)-type. Br. J. Cancer 2001, 85, 1462–1466. [Google Scholar] [CrossRef]
  378. Lugtenburg, P.J.; Löwenberg, B.; Valkema, R.; Oei, H.-Y.; Lamberts, S.W.; Eijkemans, M.J.; van Putten, W.L.; Krenning, E.P. Somatostatin receptor scintigraphy in the initial staging of low-grade non-Hodgkin’s lymphomas. J. Nucl. Med. 2001, 42, 222–229. [Google Scholar] [PubMed]
  379. Toker, C. Trabecular carcinoma of the skin. Arch. Dermatol. 1972, 105, 107–110. [Google Scholar] [CrossRef] [PubMed]
  380. Becker, J.C.; Stang, A.; DeCaprio, J.A.; Cerroni, L.; Lebbé, C.; Veness, M.; Nghiem, P. Merkel cell carcinoma. Nat. Rev. Dis. Primers 2017, 3, 17077. [Google Scholar] [CrossRef]
  381. Apraiz, A.; Benedicto, A.; Marquez, J.; Agüera-Lorente, A.; Asumendi, A.; Olaso, E.; Arteta, B. Innate Lymphoid Cells in the Malignant Melanoma Microenvironment. Cancers 2020, 12, 3177. [Google Scholar] [CrossRef] [PubMed]
  382. Reynaert, H.; Colle, I. Treatment of Advanced Hepatocellular Carcinoma with Somatostatin Analogues: A Review of the Literature. Int. J. Mol. Sci. 2019, 20, 4811. [Google Scholar] [CrossRef] [PubMed]
  383. Midorikawa, Y.; Takayama, T.; Shimada, K.; Nakayama, H.; Higaki, T.; Moriguchi, M.; Nara, S.; Tsuji, S.; Tanaka, M. Marginal survival benefit in the treatment of early hepatocellular carcinoma. J. Hepatol. 2013, 58, 306–311. [Google Scholar] [CrossRef]
  384. Midorikawa, Y.; Takayama, T.; Higaki, T.; Nakayama, H.; Yamamoto, M.; Ariizumi, S.; Shimada, K.; Kokudo, N.; Tsuji, S.; Tsuchiya, K.; et al. Early hepatocellular carcinoma as a signaling lesion for subsequent malignancy. Jpn. J. Clin. Oncol. 2016, 46, 1102–1107. [Google Scholar] [CrossRef]
  385. Ebisawa, K.; Midorikawa, Y.; Higaki, T.; Nakayama, H.; Tsuji, S.; Nishimaki, H.; Haradome, H.; Abe, O.; Sugitani, M.; Moriyama, M.; et al. Natural history of nonenhancing lesions incidentally detected during the diagnosis of hepatocellular carcinoma. Surgery 2016, 160, 654–660. [Google Scholar] [CrossRef]
  386. Yagi, R.; Midorikawa, Y.; Moriguchi, M.; Nakayama, H.; Aramaki, O.; Yamazaki, S.; Higaki, T.; Takayama, T. Liver resection for recurrent hepatocellular carcinoma to improve survivability: A proposal of indication criteria. Surgery 2018, 163, 1250–1256. [Google Scholar] [CrossRef]
  387. Shi, Y.; Taherifard, E.; Saeed, A.; Saeed, A. MASLD-Related HCC: A Comprehensive Review of the Trends, Pathophysiology, Tumor Microenvironment, Surveillance, and Treatment Options. Curr. Issues Mol. Biol. 2024, 46, 5965–5983. [Google Scholar] [CrossRef]
  388. Khaznadar, F.; Khaznadar, O.; Petrovic, A.; Hefer, M.; Gjoni, F.; Gjoni, S.; Steiner, J.; Smolic, M.; Bojanic, K. MAFLD Pandemic: Updates in Pharmacotherapeutic Approach Development. Curr. Issues Mol. Biol. 2024, 46, 6300–6314. [Google Scholar] [CrossRef] [PubMed]
  389. Wiese, S.; Voiosu, A.; Hove, J.D.; Danielsen, K.V.; Voiosu, T.; Grønbaek, H.; Møller, H.J.; Genovese, F.; Reese-Petersen, A.L.; Mookerjee, R.P.; et al. Fibrogenesis and inflammation contribute to the pathogenesis of cirrhotic cardiomyopathy. Aliment. Pharmacol. Ther. 2020, 52, 340–350. [Google Scholar] [CrossRef]
  390. Choe, J.W.; Hyun, J.J.; Kim, B.; Han, K.D. Influence of Metabolic Syndrome on Cancer Risk in HBV Carriers: A Nationwide Population Based Study Using the National Health Insurance Service Database. J. Clin. Med. 2021, 10, 2401. [Google Scholar] [CrossRef] [PubMed]
  391. Jeon, H.; Kim, J.H.; Lee, S.S.; Kim, H.J.; Cha, R.R.; Cho, H.C.; Lee, J.M.; Ha, C.Y.; Kim, H.J.; Kim, T.H.; et al. Impact of acute kidney injury on survival in patients with chronic hepatitis C: A retrospective cohort study. BMC Infect. Dis. 2021, 21, 301. [Google Scholar] [CrossRef] [PubMed]
  392. Fattovich, G.; Stroffolini, T.; Zagni, I.; Donato, F. Hepatocellular carcinoma in cirrhosis: Incidence and risk factors. Gastroenterology 2004, 127, S35–S50. [Google Scholar] [CrossRef] [PubMed]
  393. Kalambokis, G.; Fotopoulos, A.; Economou, M.; Tsianos, E.V. Octreotide in the treatment of refractory ascites of cirrhosis. Scand. J. Gastroenterol. 2006, 41, 118–121. [Google Scholar] [CrossRef] [PubMed]
  394. Zhou, D.X.; Zhou, H.B.; Wang, Q.; Zou, S.S.; Wang, H.; Hu, H.P. The effectiveness of the treatment of octreotide on chylous ascites after liver cirrhosis. Dig. Dis. Sci. 2009, 54, 1783–1788. [Google Scholar] [CrossRef]
  395. Kouroumalis, E.; Skordilis, P.; Thermos, K.; Vasilaki, A.; Moschandrea, J.; Manousos, O.N. Treatment of hepatocellular carcinoma with octreotide: A randomised controlled study. Gut 1998, 42, 442–447. [Google Scholar] [CrossRef]
  396. Wu, P.; Gu, X.Y.; Jiang, Z. Efficacy of octreotide in advanced hepatocellular carcinoma: A clinical trial. Chin. J. Hepatobiliary Surg. 2001, 7, 766–768. [Google Scholar]
  397. Pan, D.Y.; Qiao, J.G.; Chen, J.W.; Huo, Y.C.; Zhou, Y.K.; Shi, H.A. Tamoxifen combined with octreotide or regular chemotherapeutic agents in treatment of primary liver cancer: A randomized controlled trial. Hepatobiliary Pancreat. Dis. Int. 2003, 2, 211–215. [Google Scholar]
  398. Yang, M.N.; Xiao, B.; Wang, X.L.; Xue, Y.P. Effects of octreotide in elderly patients with advanced primary hepatic cancer. J. Clin. Med. Pract. 2003, 4, 302–304. [Google Scholar] [CrossRef]
  399. Patsanas, T.; Kapetanos, D.; Ilias, A.; Gessiou, C.; Tzarou, V.; Kokozidis, G.; Kitis, G. Octreotide in the treatment of inoperable hepatocellular carcinoma. Ann. Gastroenterol. 2004, 17, 69–74. [Google Scholar]
  400. Treiber, G.; Wex, T.; Röcken, C.; Fostitsch, P.; Malfertheiner, P. Impact of biomarkers on disease survival and progression in patients treated with octreotide for advanced hepatocellular carcinoma. J. Cancer Res. Clin. Oncol. 2006, 132, 699–708. [Google Scholar] [CrossRef] [PubMed]
  401. Dimitroulopoulos, D.; Xinopoulos, D.; Tsamakidis, K.; Zisimopoulos, A.; Andriotis, E.; Panagiotakos, D.; Fotopoulou, A.; Chrysohoou, C.; Bazinis, A.; Daskalopoulou, D.; et al. Long acting octreotide in the treatment of advanced hepatocellular cancer and overexpression of somatostatin receptors: Randomized placebo-controlled trial. World J. Gastroenterol. 2007, 13, 3164–3170. [Google Scholar] [CrossRef] [PubMed]
  402. Ou, S.Q.; Chen, Z.Q.; Ma, Y.L. Clinical study of octreotide for advanced hepatocellular carcinoma. Hainan Med. J. 2007, 18, 19–20. [Google Scholar]
  403. Montella, L.; Addeo, R.; Caraglia, M.; Faiola, V.; Guarrasi, R.; Vincenzi, B.; Palmeri, A.; Capasso, E.; Nocera, V.; Tarantino, L.; et al. Vascular endothelial growth factor monitoring in advanced hepatocellular carcinoma patients treated with radiofrequency ablation plus octreotide: A single center experience. Oncol. Rep. 2008, 20, 385–390. [Google Scholar] [CrossRef]
  404. Shah, U.; O’Neil, B.; Allen, J.; Goldberg, R.M.; Bernard, S.; Moore, D.; Venook, A.P.; Morse, M.M. A Phase II Study of Long-Acting Octreotide in Patients with Advanced Hepatocellular Carcinoma and CLIP Score of 3 or Higher. Gastrointest. Cancer Res. 2009, 3, 45–48. [Google Scholar]
  405. Zhang, B.; Xu, F. The clinical observation of octreotide in the treatment of 45 patients with advanced primary liver carcinoma. J. Basic Clin. Oncol. 2010, 23, 52. [Google Scholar]
  406. Prete, S.D.; Montella, L.; Caraglia, M.; Maiorino, L.; Cennamo, G.; Montesarchio, V.; Piai, G.; Febbraro, A.; Tarantino, L.; Capasso, E.; et al. Sorafenib plus octreotide is an effective and safe treatment in advanced hepatocellular carcinoma: Multicenter phase II So.LAR. study. Cancer Chemother. Pharmacol. 2010, 66, 837–844. [Google Scholar] [CrossRef]
  407. Caraglia, M.; Giuberti, G.; Marra, M.; Addeo, R.; Montella, L.; Murolo, M.; Sperlongano, P.; Vincenzi, B.; Naviglio, S.; Prete, S.D.; et al. Oxidative stress and ERK1/2 phosphorylation as predictors of outcome in hepatocellular carcinoma patients treated with sorafenib plus octreotide LAR. Cell Death Dis. 2011, 2, e150. [Google Scholar] [CrossRef]
  408. Liu, Y.; Jiang, L.; Mu, Y. Somatostatin receptor subtypes 2 and 5 are associated with better survival in operable hepatitis B-related hepatocellular carcinoma following octreotide long-acting release treatment. Oncol. Lett. 2013, 6, 821–828. [Google Scholar] [CrossRef] [PubMed]
  409. Tong, H.; Wei, B.; Chen, S.; Xie, Y.M.; Zhang, M.G.; Zhang, L.H.; Huang, Z.Y.; Tang, C.W. Adjuvant celecoxib and lanreotide following transarterial chemoembolisation for unresectable hepatocellular carcinoma: A randomized pilot study. Oncotarget 2017, 8, 48303–48312. [Google Scholar] [CrossRef] [PubMed]
  410. Raderer, M.; Hejna, M.H.; Muller, C.; Kornek, G.V.; Kurtaran, A.; Virgolini, I.; Fiebieger, W.; Hamilton, G.; Scheithauer, W. Treatment of hepatocellular cancer with the long acting somatostatin analog lanreotide in vitro and in vivo. Int. J. Oncol. 2000, 16, 1197–1201. [Google Scholar] [CrossRef]
  411. Yuen, M.F.; Poon, R.T.; Lai, C.L.; Fan, S.T.; Lo, C.M.; Wong, K.W.; Wong, W.M.; Wong, B.C. A randomized placebo-controlled study of long-acting octreotide for the treatment of advanced hepatocellular carcinoma. Hepatology 2002, 36, 687–691. [Google Scholar] [CrossRef]
  412. Rabe, C.; Pilz, T.; Allgaier, H.P.; Halm, U.; Strasser, C.; Wettstein, M.; Sauerbruch, T.; Caselmann, W.H. Clinical outcome of a cohort of 63 patients with hepatocellular carcinoma treated with octreotide. Z. Gastroenterol. 2002, 40, 395–400. [Google Scholar] [CrossRef]
  413. Plentz, R.R.; Tillmann, H.L.; Kubicka, S.; Bleck, J.S.; Gebel, M.; Manns, M.P.; Rudolph, K.L. Hepatocellular carcinoma and octreotide: Treatment results in prospectively assigned patients with advanced tumor and cirrhosis stage. J. Gastroenterol. Hepatol. 2005, 20, 1422–1428. [Google Scholar] [CrossRef]
  414. Cebon, J.; Findlay, M.; Hargreaves, C.; Stockler, M.; Thompson, P.; Boyer, M.; Roberts, S.; Poon, A.; Scott, A.M.; Kalff, V.; et al. Somatostatin receptor expression, tumour response, and quality of life in patients with advanced hepatocellular carcinoma treated with long-acting octreotide. Br. J. Cancer 2006, 95, 853–861. [Google Scholar] [CrossRef]
  415. Verset, G.; Verslype, C.; Reynaert, H.; Borbath, I.; Langlet, P.; Vandebroek, A.; Peeters, M.; Houbiers, G.; Francque, S.; Arvanitakis, M.; et al. Efficacy of the combination of long-acting release octreotide and tamoxifen in patients with advanced hepatocellular carcinoma: A randomised multicentre phase III study. Br. J. Cancer 2007, 97, 582–588. [Google Scholar] [CrossRef] [PubMed]
  416. Becker, G.; Allgaier, H.P.; Olschewski, M.; Zahringer, A.; Blum, H.E.; Group, H.S. Long-acting octreotide versus placebo for treatment of advanced HCC: A randomized controlled double-blind study. Hepatology 2007, 45, 9–15. [Google Scholar] [CrossRef]
  417. Barbare, J.C.; Bouche, O.; Bonnetain, F.; Dahan, L.; Lombard-Bohas, C.; Faroux, R.; Raoul, J.L.; Cattan, S.; Lemoine, A.; Blanc, J.F.; et al. Treatment of advanced hepatocellular carcinoma with long-acting octreotide: A phase III multicentre, randomised, double blind placebo-controlled study. Eur. J. Cancer 2009, 45, 1788–1797. [Google Scholar] [CrossRef]
  418. Lanreotide. In Drugs and Lactation Database (LactMed®); National Institute of Child Health and Human Development: Bethesda, MD, USA, 2006.
  419. Paulson, S.; Ray, D.; Aranha, S.; Scales, A.; Wang, Y.; Liu, E. Lanreotide Depot to Treat Gastroenteropancreatic Neuroendocrine Tumors in a US Community Oncology Setting: A Prospective, Observational Study. Oncol. Ther. 2022, 10, 463–479. [Google Scholar] [CrossRef] [PubMed]
  420. Lepage, C.; Phelip, J.M.; Lievre, A.; Le-Malicot, K.; Dahan, L.; Tougeron, D.; Toumpanakis, C.; Di-Fiore, F.; Lombard-Bohas, C.; Borbath, I.; et al. Lanreotide as maintenance therapy after first-line treatment in patients with non-resectable duodeno-pancreatic neuroendocrine tumours: An international double-blind, placebo-controlled randomised phase II trial—Prodige 31 REMINET: An FFCD study. Eur. J. Cancer 2022, 175, 31–40. [Google Scholar] [CrossRef] [PubMed]
  421. Sills, E.S.; Wood, S.H.; Tan, S.L.; Ibach, D.M. Neuroendocrine tumor chromogranin A response following synthetic somatostatin analog (lanreotide): Early observations from an isolated duodenal neoplasm. Neuro Endocrinol. Lett. 2023, 44, 265–269. [Google Scholar]
  422. Hautefeuille, V.; Walter, T.; Do Cao, C.; Coriat, R.; Dominguez, S.; Mineur, L.; Cadiot, G.; Terrebonne, E.; Sobhani, I.; Gueguen, D.; et al. OPERA: Perception of information in patients with gastroenteropancreatic neuroendocrine tumors on lanreotide autogel. Eur. J. Endocrinol. 2023, 189, 281–289. [Google Scholar] [CrossRef]
  423. O’Toole, D.; Kunz, P.L.; Webb, S.M.; Goldstein, G.; Khawaja, S.; McDonnell, M.; Boiziau, S.; Gueguen, D.; Houchard, A.; Ribeiro-Oliveira, A., Jr.; et al. PRESTO 2: An International Survey to Evaluate Patients’ Injection Experiences with the Latest Devices/Formulations of Long-Acting Somatostatin Analog Therapies for Neuroendocrine Tumors or Acromegaly. Adv. Ther. 2023, 40, 671–690. [Google Scholar] [CrossRef]
  424. Hernando, J.; Kolarova, T.; Verslype, C.; Kaltsas, G.; Houchard, A.; Gueguen, D.; De Herder, W.W. Satisfaction with injection experience of patients with neuroendocrine tumors enrolled on lanreotide autogel patient support programs: Results from the international HomeLAN survey. J. Neuroendocrinol. 2023, 35, e13281. [Google Scholar] [CrossRef]
  425. Raj, N.; Cruz, E.; O’Shaughnessy, S.; Calderon, C.; Chou, J.F.; Capanu, M.; Heffernan, O.; DeMore, A.; Punn, S.; Le, T.; et al. A Randomized Trial Evaluating Patient Experience and Preference between Octreotide Long-Acting Release and Lanreotide for Treatment of Well-Differentiated Neuroendocrine Tumors. JCO Oncol. Pract. 2022, 18, e1533–e1541. [Google Scholar] [CrossRef] [PubMed]
  426. Pavel, M.; Lahner, H.; Hörsch, D.; Rinke, A.; Denecke, T.; Koch, A.; Regnault, B.; Helbig, D.; Hoffmanns, P.; Raderer, M. Combined Lanreotide Autogel and Temozolomide Treatment of Progressive Pancreatic and Intestinal Neuroendocrine Tumors: The Phase II SONNET Study. Oncologist 2024, 29, e643–e654. [Google Scholar] [CrossRef] [PubMed]
  427. Sood, A.; Munir, M.; Syed, O.; Mehta, V.; Kaur, R.; Kumar, A.; Sridhar, A.; Sood, A.; Gupta, R. An update on the safety of lanreotide autogel for the treatment of patients with neuroendocrine tumors. Expert Opin. Drug Saf. 2024, 23, 949–957. [Google Scholar] [CrossRef]
  428. Shiraishi, K.; Akai, T.; Tomita, T.; Hayashi, R.; Minamisaka, T.; Kuroda, S. Pituitary apoplexy in endocrinologically silent adenoma during somatostatin analog administration for pancreatic neuroendocrine tumor: A case report. Neuropathology 2024, 44, 247–251. [Google Scholar] [CrossRef]
  429. Pasireotide. In Drugs and Lactation Database (LactMed®); National Institute of Child Health and Human Development: Bethesda, MD, USA, 2006.
  430. Bruns, C.; Lewis, I.; Briner, U.; Meno-Tetang, G.; Weckbecker, G. SOM230: A novel somatostatin peptidomimetic with broad somatotropin release inhibiting factor (SRIF) receptor binding and a unique antisecretory profile. Eur. J. Endocrinol. 2002, 146, 707–716. [Google Scholar] [CrossRef]
  431. Hofland, L.J.; Lamberts, S.W. The pathophysiological consequences of somatostatin receptor internalization and resistance. Endocr. Rev. 2003, 24, 28–47. [Google Scholar] [CrossRef] [PubMed]
  432. Ronga, G.; Salerno, G.; Procaccini, E.; Mauro, L.; Annovazzi, A.; Barone, R.; Mellozzi, M.; Tamburrano, G.; Signore, A. 111In-octreotide scintigraphy in metastatic medullary thyroid carcinoma before and after octreotide therapy: In vivo evidence of the possible down-regulation of somatostatin receptors. Q. J. Nucl. Med. 1995, 39, 134–136. [Google Scholar] [PubMed]
  433. Li, M.; Li, W.; Kim, H.J.; Yao, Q.; Chen, C.; Fisher, W.E. Characterization of somatostatin receptor expression in human pancreatic cancer using real-time RT-PCR. J. Surg. Res. 2004, 119, 130–137. [Google Scholar] [CrossRef]
  434. Schmid, H.A.; Schoeffter, P. Functional activity of the multiligand analog SOM230 at human recombinant somatostatin receptor subtypes supports its usefulness in neuroendocrine tumors. Neuroendocrinology 2004, 80 (Suppl. S1), 47–50. [Google Scholar] [CrossRef] [PubMed]
  435. Allen, P.J.; Gönen, M.; Brennan, M.F.; Bucknor, A.A.; Robinson, L.M.; Pappas, M.M.; Carlucci, K.E.; D’Angelica, M.I.; DeMatteo, R.P.; Kingham, T.P.; et al. Pasireotide for postoperative pancreatic fistula. N. Engl. J. Med. 2014, 370, 2014–2022. [Google Scholar] [CrossRef]
  436. Efstathiadou, Z.A.; Divaris, E.; Michou, A.; Kyriakopoulos, G.; Kita, M.D. Complete and sustained remission of hypercortisolism with pasireotide treatment of an adrenocorticotropic hormone (ACTH)-secreting thoracic neuroendocrine tumor: An n-of-1 trial. Endocr. J. 2023, 70, 229–232. [Google Scholar] [CrossRef]
  437. Sanoff, H.K.; Kim, R.; Ivanova, A.; Alistar, A.; McRee, A.J.; O’Neil, B.H. Everolimus and pasireotide for advanced and metastatic hepatocellular carcinoma. Investig. New Drugs 2015, 33, 505–509. [Google Scholar] [CrossRef] [PubMed]
  438. Feun, L.G.; Wangpaichitr, M.; Li, Y.Y.; Kwon, D.; Richman, S.P.; Hosein, P.J.; Savaraj, N. Phase II trial of SOM230 (pasireotide LAR) in patients with unresectable hepatocellular carcinoma. J. Hepatocell. Carcinoma 2018, 5, 9–15. [Google Scholar] [CrossRef] [PubMed]
  439. Wagner, H.N., Jr.; Burns, H.D.; Dannals, R.F.; Wong, D.F.; Langstrom, B.; Duelfer, T.; Frost, J.J.; Ravert, H.T.; Links, J.M.; Rosenbloom, S.B.; et al. Imaging dopamine receptors in the human brain by positron tomography. Science 1983, 221, 1264–1266. [Google Scholar] [CrossRef]
  440. Pishdad, R.; Treglia, G.; Mehta, A.; Santhanam, P. Somatostatin receptor imaging of thyroid tissue and differentiated thyroid cancer using gallium-68-labeled radiotracers—A review of clinical studies. Endocrine 2024, 85, 566–575. [Google Scholar] [CrossRef]
  441. Sakellis, C.; Jacene, H.A. Neuroendocrine Tumors: Diagnostics. PET Clin. 2024, 19, 325–339. [Google Scholar] [CrossRef] [PubMed]
  442. Mallak, N.; O’Brien, S.R.; Pryma, D.A.; Mittra, E. Theranostics in Neuroendocrine Tumors. Cancer J. 2024, 30, 185–193. [Google Scholar] [CrossRef] [PubMed]
  443. Park, S.; Parihar, A.S.; Bodei, L.; Hope, T.A.; Mallak, N.; Millo, C.; Prasad, K.; Wilson, D.; Zukotynski, K.; Mittra, E. Somatostatin Receptor Imaging and Theranostics: Current Practice and Future Prospects. J. Nucl. Med. 2021, 62, 1323–1329. [Google Scholar] [CrossRef]
  444. Bass, R.T.; Buckwalter, B.L.; Patel, B.P.; Pausch, M.H.; Price, L.A.; Strnad, J.; Hadcock, J.R. Identification and characterization of novel somatostatin antagonists. Mol. Pharmacol. 1996, 50, 709–715. [Google Scholar] [PubMed]
  445. Ginj, M.; Zhang, H.; Waser, B.; Cescato, R.; Wild, D.; Wang, X.; Erchegyi, J.; Rivier, J.; Mäcke, H.R.; Reubi, J.C. Radiolabeled somatostatin receptor antagonists are preferable to agonists for in vivo peptide receptor targeting of tumors. Proc. Natl. Acad. Sci. USA 2006, 103, 16436–16441. [Google Scholar] [CrossRef]
  446. Wild, D.; Fani, M.; Behe, M.; Brink, I.; Rivier, J.E.; Reubi, J.C.; Maecke, H.R.; Weber, W.A. First clinical evidence that imaging with somatostatin receptor antagonists is feasible. J. Nucl. Med. 2011, 52, 1412–1417. [Google Scholar] [CrossRef]
  447. Fani, M.; Braun, F.; Waser, B.; Beetschen, K.; Cescato, R.; Erchegyi, J.; Rivier, J.E.; Weber, W.A.; Maecke, H.R.; Reubi, J.C. Unexpected sensitivity of sst2 antagonists to N-terminal radiometal modifications. J. Nucl. Med. 2012, 53, 1481–1489. [Google Scholar] [CrossRef]
  448. Modarai, S.R.; Opdenaker, L.M.; Viswanathan, V.; Fields, J.Z.; Boman, B.M. Somatostatin signaling via SSTR1 contributes to the quiescence of colon cancer stem cells. BMC Cancer 2016, 16, 941. [Google Scholar] [CrossRef]
  449. Kasprzak, A.; Geltz, A. The State-of-the-Art Mechanisms and Antitumor Effects of Somatostatin in Colorectal Cancer: A Review. Biomedicines 2024, 12, 578. [Google Scholar] [CrossRef]
  450. Kimura, N.; Hayafuji, C.; Kimura, N. Characterization of 17-beta-estradiol-dependent and -independent somatostatin receptor subtypes in rat anterior pituitary. J. Biol. Chem. 1989, 264, 7033–7040. [Google Scholar] [CrossRef] [PubMed]
  451. Seredycz, L.I.; Lautt, W.W. Hemorrhage results in hepatic insulin-sensitizing substance-dependent insulin resistance mediated by somatostatin in rats. Neuroendocrinology 2006, 84, 94–102. [Google Scholar] [CrossRef]
  452. Ikeda, H.; Kotani, A.; Koshikawa, N.; Cools, A.R. Somatostatin receptors in the nucleus accumbens modulate dopamine-dependent but not acetylcholine-dependent turning behaviour of rats. Neuroscience 2009, 159, 974–981. [Google Scholar] [CrossRef]
  453. Ionov, I.D.; Pushinskaya, I.I. Somatostatin antagonist induces catalepsy in the aged rat. Psychopharmacology 2013, 227, 273–276. [Google Scholar] [CrossRef] [PubMed]
  454. Jiang, J.; Peng, Y.; He, Z.; Wei, L.; Jin, W.; Wang, X.; Chang, M. Intrahippocampal injection of Cortistatin-14 impairs recognition memory consolidation in mice through activation of sst2, ghrelin and GABAA/B receptors. Brain Res. 2017, 1666, 38–47. [Google Scholar] [CrossRef]
  455. Ionov, I.D.; Pushinskaya, I.I.; Gorev, N.P.; Frenkel, D.D. Cyclosomatostatin- and haloperidol-induced catalepsy in Wistar rats: Differential responsiveness to sleep deprivation. Neurosci. Lett. 2018, 684, 72–77. [Google Scholar] [CrossRef] [PubMed]
  456. Ionov, I.D.; Pushinskaya, I.I.; Roslavtseva, L.A.; Severtsev, N.N. Brain sites mediating cyclosomatostatin-induced catalepsy in Wistar rats: A specific role for the nigrostriatal system and locus coeruleus. Brain Res. 2018, 1691, 26–33. [Google Scholar] [CrossRef]
  457. Ionov, I.D.; Pushinskaya, I.I.; Gorev, N.P.; Frenkel, D.D. Cyclosomatostatin-induced catalepsy in aged rats: Specific change of brain c-Fos protein expression in the lateral entorhinal cortex. Brain Res. Bull. 2020, 159, 79–86. [Google Scholar] [CrossRef]
  458. Sekiya, H.; Yokota, N.; Takemi, S.; Nakayama, K.; Okada, H.; Sakai, T.; Sakata, I. The inhibitory effect of somatostatin on gastric motility in Suncus murinus. J. Smooth Muscle Res. 2020, 56, 69–81. [Google Scholar] [CrossRef]
  459. Ionov, I.D.; Pushinskaya, I.I.; Gorev, N.P.; Frenkel, D.D.; Severtsev, N.N. Anticataleptic activity of nicotine in rats: Involvement of the lateral entorhinal cortex. Psychopharmacology 2021, 238, 2471–2483. [Google Scholar] [CrossRef]
  460. Croze, M.L.; Flisher, M.F.; Guillaume, A.; Tremblay, C.; Noguchi, G.M.; Granziera, S.; Vivot, K.; Castillo, V.C.; Campbell, S.A.; Ghislain, J.; et al. Free fatty acid receptor 4 inhibitory signaling in delta cells regulates islet hormone secretion in mice. Mol. Metab. 2021, 45, 101166. [Google Scholar] [CrossRef] [PubMed]
  461. Hernandez-Unzueta, I.; Benedicto, A.; Telleria, U.; Sanz, E.; Márquez, J. Improving the Antitumor Effect of Chemotherapy with Ocoxin as a Novel Adjuvant Agent to Treat Prostate Cancer. Nutrients 2023, 15, 2536. [Google Scholar] [CrossRef] [PubMed]
  462. Shalhout, S.Z.; Miller, D.M.; Emerick, K.S.; Kaufman, H.L. Therapy with oncolytic viruses: Progress and challenges. Nat. Rev. Clin. Oncol. 2023, 20, 160–177. [Google Scholar] [CrossRef]
  463. Kaur, R.; Bhardwaj, A.; Gupta, S. Cancer treatment therapies: Traditional to modern approaches to combat cancers. Mol. Biol. Rep. 2023, 50, 9663–9676. [Google Scholar] [CrossRef]
  464. Sprecher, U.; Mohr, P.; Martin, R.E.; Maerki, H.P.; Sanchez, R.A.; Binggeli, A.; Künnecke, B.; Christ, A.D. Novel, non-peptidic somatostatin receptor subtype 5 antagonists improve glucose tolerance in rodents. Regul. Pept. 2010, 159, 19–27. [Google Scholar] [CrossRef]
  465. Hirose, H.; Yamasaki, T.; Ogino, M.; Mizojiri, R.; Tamura-Okano, Y.; Yashiro, H.; Muraki, Y.; Nakano, Y.; Sugama, J.; Hata, A.; et al. Discovery of novel 5-oxa-2,6-diazaspiro[3.4]oct-6-ene derivatives as potent, selective, and orally available somatostatin receptor subtype 5 (SSTR5) antagonists for treatment of type 2 diabetes mellitus. Bioorganic Med. Chem. 2017, 25, 4175–4193. [Google Scholar] [CrossRef]
  466. Feniuk, W.; Jarvie, E.; Luo, J.; Humphrey, P.P. Selective somatostatin sst2 receptor blockade with the novel cyclic octapeptide, CYN-154806. Neuropharmacology 2000, 39, 1443–1450. [Google Scholar] [CrossRef] [PubMed]
  467. Takeuchi, K.; Endoh, T.; Hayashi, S.; Aihara, T. Activation of Muscarinic Acetylcholine Receptor Subtype 4 Is Essential for Cholinergic Stimulation of Gastric Acid Secretion: Relation to D Cell/Somatostatin. Front. Pharmacol. 2016, 7, 278. [Google Scholar] [CrossRef]
  468. He, S.; Ye, Z.; Truong, Q.; Shah, S.; Du, W.; Guo, L.; Dobbelaar, P.H.; Lai, Z.; Liu, J.; Jian, T.; et al. The Discovery of MK-4256, a Potent SSTR3 Antagonist as a Potential Treatment of Type 2 Diabetes. ACS Med. Chem. Lett. 2012, 3, 484–489. [Google Scholar] [CrossRef]
  469. Hocart, S.J.; Jain, R.; Murphy, W.A.; Taylor, J.E.; Coy, D.H. Highly potent cyclic disulfide antagonists of somatostatin. J. Med. Chem. 1999, 42, 1863–1871. [Google Scholar] [CrossRef]
  470. Rossowski, W.J.; Cheng, B.L.; Taylor, J.E.; Datta, R.; Coy, D.H. Human urotensin II-induced aorta ring contractions are mediated by protein kinase C, tyrosine kinases and Rho-kinase: Inhibition by somatostatin receptor antagonists. Eur. J. Pharmacol. 2002, 438, 159–170. [Google Scholar] [CrossRef] [PubMed]
  471. Feniuk, W.; Dimech, J.; Jarvie, E.M.; Humphrey, P.P. Further evidence from functional studies for somatostatin receptor heterogeneity in guinea-pig isolated ileum, vas deferens and right atrium. Br. J. Pharmacol. 1995, 115, 975–980. [Google Scholar] [CrossRef]
  472. Liu, W.; Shao, P.P.; Liang, G.B.; Bawiec, J.; He, J.; Aster, S.D.; Wu, M.; Chicchi, G.; Wang, J.; Tsao, K.L.; et al. Discovery and Pharmacology of a Novel Somatostatin Subtype 5 (SSTR5) Antagonist: Synergy with DPP-4 Inhibition. ACS Med. Chem. Lett. 2018, 9, 1082–1087. [Google Scholar] [CrossRef]
  473. Saveanu, A.; Lavaque, E.; Gunz, G.; Barlier, A.; Kim, S.; Taylor, J.E.; Culler, M.D.; Enjalbert, A.; Jaquet, P. Demonstration of enhanced potency of a chimeric somatostatin-dopamine molecule, BIM-23A387, in suppressing growth hormone and prolactin secretion from human pituitary somatotroph adenoma cells. J. Clin. Endocrinol. Metab. 2002, 87, 5545–5552. [Google Scholar] [CrossRef] [PubMed]
  474. Jaquet, P.; Gunz, G.; Saveanu, A.; Dufour, H.; Taylor, J.; Dong, J.; Kim, S.; Moreau, J.P.; Enjalbert, A.; Culler, M.D. Efficacy of chimeric molecules directed towards multiple somatostatin and dopamine receptors on inhibition of GH and prolactin secretion from GH-secreting pituitary adenomas classified as partially responsive to somatostatin analog therapy. Eur. J. Endocrinol. 2005, 153, 135–141. [Google Scholar] [CrossRef] [PubMed]
  475. Rocheville, M.; Lange, D.C.; Kumar, U.; Patel, S.C.; Patel, R.C.; Patel, Y.C. Receptors for dopamine and somatostatin: Formation of hetero-oligomers with enhanced functional activity. Science 2000, 288, 154–157. [Google Scholar] [CrossRef]
  476. Miki, Y. Hormone-Dependent Cancers: New Aspects on Biochemistry and Molecular Pathology. Int. J. Mol. Sci. 2023, 24, 10830. [Google Scholar] [CrossRef]
  477. Ulm, M.; Ramesh, A.V.; McNamara, K.M.; Ponnusamy, S.; Sasano, H.; Narayanan, R. Therapeutic advances in hormone-dependent cancers: Focus on prostate, breast and ovarian cancers. Endocr. Connect. 2019, 8, R10–R26. [Google Scholar] [CrossRef] [PubMed]
  478. Emons, G. Hormone-Dependent Cancers: Molecular Mechanisms and Therapeutical Implications. Cells 2023, 12, 110. [Google Scholar] [CrossRef]
  479. Stueven, A.K.; Kayser, A.; Wetz, C.; Amthauer, H.; Wree, A.; Tacke, F.; Wiedenmann, B.; Roderburg, C.; Jann, H. Somatostatin Analogues in the Treatment of Neuroendocrine Tumors: Past, Present and Future. Int. J. Mol. Sci. 2019, 20, 3049. [Google Scholar] [CrossRef]
  480. Kong, G.; Callahan, J.; Hofman, M.S.; Pattison, D.A.; Akhurst, T.; Michael, M.; Eu, P.; Hicks, R.J. High clinical and morphologic response using (90)Y-DOTA-octreotate sequenced with (177)Lu-DOTA-octreotate induction peptide receptor chemoradionuclide therapy (PRCRT) for bulky neuroendocrine tumours. Eur. J. Nucl. Med. Mol. Imaging 2017, 44, 476–489. [Google Scholar] [CrossRef] [PubMed]
  481. Kumar, U. Somatostatin and Somatostatin Receptors in Tumour Biology. Int. J. Mol. Sci. 2023, 25, 436. [Google Scholar] [CrossRef] [PubMed]
  482. Angelousi, A.; Koumarianou, A.; Chatzellis, E.; Kaltsas, G. Resistance of neuroendocrine tumours to somatostatin analogs. Expert Rev. Endocrinol. Metab. 2023, 18, 33–52. [Google Scholar] [CrossRef]
  483. Siddiqui, Z.; Marginean, H.; Leung, M.; Asmis, T.; Vickers, M.; Goodwin, R. Real world use of lanreotide in neuroendocrine tumors. J. Gastrointest. Oncol. 2023, 14, 1488–1495. [Google Scholar] [CrossRef]
  484. Hatzoglou, A.; Bakogeorgou, E.; Kampa, M.; Panagiotou, S.; Martin, P.M.; Loukas, S.; Castanas, E. Somatostatin and opioid receptors in mammary tissue. Role in cancer cell growth. Adv. Exp. Med. Biol. 2000, 480, 55–63. [Google Scholar] [CrossRef]
  485. Drewe, J.; Sieber, C.C.; Mottet, C.; Wullschleger, C.; Larsen, F.; Beglinger, C. Dose-dependent gastrointestinal effects of the somatostatin analog lanreotide in healthy volunteers. Clin. Pharmacol. Ther. 1999, 65, 413–419. [Google Scholar] [CrossRef]
  486. Juliana, C.A.; Chai, J.; Arroyo, P.; Rico-Bautista, E.; Betz, S.F.; De León, D.D. A selective nonpeptide somatostatin receptor 5 agonist effectively decreases insulin secretion in hyperinsulinism. J. Biol. Chem. 2023, 299, 104816. [Google Scholar] [CrossRef]
  487. George, J.; Ramage, J.; White, B.; Srirajaskanthan, R. The role of serotonin inhibition within the treatment of carcinoid syndrome. Endocr. Oncol. 2023, 3, e220077. [Google Scholar] [CrossRef] [PubMed]
  488. Melhorn, P.; Kretschmer-Chott, E.; Wolff, L.; Haug, A.; Mazal, P.; Raderer, M.; Kiesewetter, B. Treatment patterns and oncological outcome of patients with advanced small intestinal neuroendocrine tumors: Real-world data from the Medical University of Vienna. Ther. Adv. Med. Oncol. 2022, 14, 17588359221138389. [Google Scholar] [CrossRef]
  489. Caruntu, A.; Moraru, L.; Surcel, M.; Munteanu, A.; Costache, D.O.; Tanase, C.; Constantin, C.; Scheau, C.; Neagu, M.; Caruntu, C. Persistent Changes of Peripheral Blood Lymphocyte Subsets in Patients with Oral Squamous Cell Carcinoma. Healthcare 2022, 10, 342. [Google Scholar] [CrossRef]
  490. König, R.; Zhou, W. Signal transduction in T helper cells: CD4 coreceptors exert complex regulatory effects on T cell activation and function. Curr. Issues Mol. Biol. 2004, 6, 1–15. [Google Scholar] [CrossRef]
  491. Lee, H.W.; Park, C.; Joung, J.G.; Kang, M.; Chung, Y.S.; Oh, W.J.; Yeom, S.Y.; Park, W.Y.; Kim, T.J.; Seo, S.I. Renal Cell Carcinoma-Infiltrating CD3low Vγ9Vδ1 T Cells Represent Potentially Novel Anti-Tumor Immune Players. Curr. Issues Mol. Biol. 2021, 43, 226–239. [Google Scholar] [CrossRef] [PubMed]
  492. Machado-Alba, J.E.; Machado-Duque, M.E.; Gaviria-Mendoza, A.; Arsof-Saab, I.N.; Castellanos-Moreno, C.A.; Botero, L.; Triana, L. Prescription patterns of somatostatin analogs in patients with acromegaly and neuroendocrine tumors. J. Endocrinol. Investig. 2023, 46, 27–35. [Google Scholar] [CrossRef] [PubMed]
  493. Jones, S.L.; Patchett, S.; Anderson, J.V.; Farthing, M.J.; Besser, G.M.; Wass, J.A. Prevalence of Helicobacter pylori in acromegalic patients during treatment with octreotide. Clin. Endocrinol. 1995, 43, 683–687. [Google Scholar] [CrossRef]
  494. Anderson, J.V.; Catnach, S.; Lowe, D.G.; Fairclough, P.D.; Besser, G.M.; Wass, J.A. Prevalence of gastritis in patients with acromegaly: Untreated and during treatment with octreotide. Clin. Endocrinol. 1992, 37, 227–232. [Google Scholar] [CrossRef]
  495. Xu, W.F.; Wang, Y.; Huang, H.; Wu, J.W.; Che, Y.; Ding, C.J.; Zhang, Q.; Cao, W.L.; Cao, L.J. Octreotide-based therapies effectively protect mice from acute and chronic gastritis. Eur. J. Pharmacol. 2022, 928, 174976. [Google Scholar] [CrossRef] [PubMed]
  496. Thán, M.; Németh, J.; Szilvássy, Z.; Pintér, E.; Helyes, Z.; Szolcsányi, J. Systemic anti-inflammatory effect of somatostatin released from capsaicin-sensitive vagal and sciatic sensory fibres of the rat and guinea-pig. Eur. J. Pharmacol. 2000, 399, 251–258. [Google Scholar] [CrossRef]
  497. Szolcsányi, J.; Barthó, L. Capsaicin-sensitive afferents and their role in gastroprotection: An update. J. Physiol. Paris. 2001, 95, 181–188. [Google Scholar] [CrossRef]
  498. Periferakis, A.T.; Periferakis, A.; Periferakis, K.; Caruntu, A.; Badarau, I.A.; Savulescu-Fiedler, I.; Scheau, C.; Caruntu, C. Antimicrobial Properties of Capsaicin: Available Data and Future Research Perspectives. Nutrients 2023, 15, 4097. [Google Scholar] [CrossRef]
  499. Petran, E.M.; Periferakis, A.; Troumpata, L.; Periferakis, A.-T.; Scheau, A.-E.; Badarau, I.A.; Periferakis, K.; Caruntu, A.; Savulescu-Fiedler, I.; Sima, R.-M.; et al. Capsaicin: Emerging Pharmacological and Therapeutic Insights. Curr. Issues Mol. Biol. 2024, 46, 7895–7943. [Google Scholar] [CrossRef]
  500. Jancsó, G.; Király, E.; Such, G.; Joó, F.; Nagy, A. Neurotoxic effect of capsaicin in mammals. Acta Physiol. Hung. 1987, 69, 295–313. [Google Scholar] [PubMed]
  501. Ritter, S.; Dinh, T.T. Capsaicin-induced neuronal degeneration in the brain and retina of preweanling rats. J. Comp. Neurol. 1990, 296, 447–461. [Google Scholar] [CrossRef] [PubMed]
  502. Satyanarayana, M.N. Capsaicin and gastric ulcers. Crit. Rev. Food Sci. Nutr. 2006, 46, 275–328. [Google Scholar] [CrossRef] [PubMed]
  503. Georgescu, S.R.; Sârbu, M.I.; Matei, C.; Ilie, M.A.; Caruntu, C.; Constantin, C.; Neagu, M.; Tampa, M. Capsaicin: Friend or Foe in Skin Cancer and Other Related Malignancies? Nutrients 2017, 9, 1365. [Google Scholar] [CrossRef]
  504. Scheau, C.; Mihai, L.; Bădărău, I.; Caruntu, C. Emerging applications of some important natural compounds in the field of oncology. Farmacia 2020, 68, 984–991. [Google Scholar] [CrossRef]
  505. Dumitrache, M.D.; Jieanu, A.S.; Scheau, C.; Badarau, I.A.; Popescu, G.D.A.; Caruntu, A.; Costache, D.O.; Costache, R.S.; Constantin, C.; Neagu, M.; et al. Comparative effects of capsaicin in chronic obstructive pulmonary disease and asthma (Review). Exp. Ther. Med. 2021, 22, 917. [Google Scholar] [CrossRef]
  506. Popescu, G.D.A.; Scheau, C.; Badarau, I.A.; Dumitrache, M.D.; Caruntu, A.; Scheau, A.E.; Costache, D.O.; Costache, R.S.; Constantin, C.; Neagu, M.; et al. The Effects of Capsaicin on Gastrointestinal Cancers. Molecules 2020, 26, 94. [Google Scholar] [CrossRef]
  507. Scheau, C.; Badarau, I.A.; Caruntu, C.; Mihai, G.L.; Didilescu, A.C.; Constantin, C.; Neagu, M. Capsaicin: Effects on the Pathogenesis of Hepatocellular Carcinoma. Molecules 2019, 24, 2350. [Google Scholar] [CrossRef]
  508. Srinivasan, K. Biological Activities of Red Pepper (Capsicum annuum) and Its Pungent Principle Capsaicin: A Review. Crit. Rev. Food Sci. Nutr. 2016, 56, 1488–1500. [Google Scholar] [CrossRef]
  509. Bencze, N.; Schvarcz, C.; Kriszta, G.; Danics, L.; Szőke, É.; Balogh, P.; Szállási, Á.; Hamar, P.; Helyes, Z.; Botz, B. Desensitization of Capsaicin-Sensitive Afferents Accelerates Early Tumor Growth via Increased Vascular Leakage in a Murine Model of Triple Negative Breast Cancer. Front. Oncol. 2021, 11, 685297. [Google Scholar] [CrossRef]
  510. Zhang, X.; Yang, C.; Zhou, J.; Huo, M. Somatostatin Receptor-Mediated Tumor-Targeting Nanocarriers Based on Octreotide-PEG Conjugated Nanographene Oxide for Combined Chemo and Photothermal Therapy. Small 2016, 12, 3578–3590. [Google Scholar] [CrossRef]
  511. Aiello, P.; Consalvi, S.; Poce, G.; Raguzzini, A.; Toti, E.; Palmery, M.; Biava, M.; Bernardi, M.; Kamal, M.A.; Perry, G.; et al. Dietary flavonoids: Nano delivery and nanoparticles for cancer therapy. Semin. Cancer Biol. 2021, 69, 150–165. [Google Scholar] [CrossRef]
  512. Khan, H.; Ullah, H.; Martorell, M.; Valdes, S.E.; Belwal, T.; Tejada, S.; Sureda, A.; Kamal, M.A. Flavonoids nanoparticles in cancer: Treatment, prevention and clinical prospects. Semin. Cancer Biol. 2021, 69, 200–211. [Google Scholar] [CrossRef] [PubMed]
  513. Matei, A.-M.; Caruntu, C.; Tampa, M.; Georgescu, S.R.; Matei, C.; Constantin, M.M.; Constantin, T.V.; Calina, D.; Ciubotaru, D.A.; Badarau, I.A.; et al. Applications of Nanosized-Lipid-Based Drug Delivery Systems in Wound Care. Appl. Sci. 2021, 11, 4915. [Google Scholar] [CrossRef]
  514. Li, X.; Chen, L.; Luan, S.; Zhou, J.; Xiao, X.; Yang, Y.; Mao, C.; Fang, P.; Chen, L.; Zeng, X.; et al. The development and progress of nanomedicine for esophageal cancer diagnosis and treatment. Semin. Cancer Biol. 2022, 86, 873–885. [Google Scholar] [CrossRef] [PubMed]
  515. Allen, T.M.; Cullis, P.R. Liposomal drug delivery systems: From concept to clinical applications. Adv. Drug Deliv. Rev. 2013, 65, 36–48. [Google Scholar] [CrossRef] [PubMed]
  516. Alavi, M.; Hamidi, M. Passive and active targeting in cancer therapy by liposomes and lipid nanoparticles. Drug Metab. Pers. Ther. 2019, 34, 20180032. [Google Scholar] [CrossRef]
  517. Chiu, H.Y.; Hsieh, Y.J.; Tsai, P.S. Acupuncture to Reduce Sleep Disturbances in Perimenopausal and Postmenopausal Women: A Systematic Review and Meta-analysis. Obstet. Gynecol. 2016, 127, 507–515. [Google Scholar] [CrossRef]
  518. Fu, H.; Sun, J.; Tan, Y.; Zhou, H.; Xu, W.; Zhou, J.; Chen, D.; Zhang, C.; Zhu, X.; Zhang, Y.; et al. Effects of acupuncture on the levels of serum estradiol and pituitary estrogen receptor beta in a rat model of induced super ovulation. Life Sci. 2018, 197, 109–113. [Google Scholar] [CrossRef]
  519. Dong, X.L.; Ran, J.K.; Zhang, H.J.; Chen, K.; Li, H.X. Acupuncture combined with medication improves endocrine hormone levels and ovarian reserve function in poor ovarian response patients undergoing in vitro fertilization-embryo transplantation. Zhen Ci Yan Jiu 2019, 44, 599–604. [Google Scholar] [CrossRef]
  520. Li, X.; Wu, Z.; Chen, Y.; Cai, R.; Wang, Z. Effect of Acupuncture on Simple Obesity and Serum Levels of Prostaglandin E and Leptin in Sprague-Dawley Rats. Comput. Math. Methods Med. 2021, 2021, 6730274. [Google Scholar] [CrossRef] [PubMed]
  521. Liu, C.; Wang, Z.; Guo, T.; Zhuang, L.; Gao, X. Effect of acupuncture on menopausal hot flushes and serum hormone levels: A systematic review and meta-analysis. Acupunct. Med. 2022, 40, 289–298. [Google Scholar] [CrossRef]
  522. Amorim, D.; Brito, I.; Caseiro, A.; Figueiredo, J.P.; Pinto, A.; Macedo, I.; Machado, J. Electroacupuncture and acupuncture in the treatment of anxiety—A double blinded randomized parallel clinical trial. Complement. Ther. Clin. Pract. 2022, 46, 101541. [Google Scholar] [CrossRef]
  523. Periferakis, K.; Periferakis, A. Treating gout caused by renal insufficiency with acupuncture and moxibustion: A case report. Rom. J. Clin. Res. 2023, 6, 36–43. [Google Scholar]
  524. O’Regan, D.; Filshie, J. Acupuncture and cancer. Auton. Neurosci. 2010, 157, 96–100. [Google Scholar] [CrossRef]
  525. Yang, J.; Wahner-Roedler, D.L.; Zhou, X.; Johnson, L.A.; Do, A.; Pachman, D.R.; Chon, T.Y.; Salinas, M.; Millstine, D.; Bauer, B.A. Acupuncture for palliative cancer pain management: Systematic review. BMJ Support. Palliat. Care 2021, 11, 264–270. [Google Scholar] [CrossRef] [PubMed]
  526. Zhang, J.; Zhang, Z.; Huang, S.; Qiu, X.; Lao, L.; Huang, Y.; Zhang, Z.J. Acupuncture for cancer-related insomnia: A systematic review and meta-analysis. Phytomedicine 2022, 102, 154160. [Google Scholar] [CrossRef] [PubMed]
  527. Zhang, X.W.; Hou, W.B.; Pu, F.L.; Wang, X.F.; Wang, Y.R.; Yang, M.; Cheng, K.; Wang, Y.; Robinson, N.; Liu, J.P. Acupuncture for cancer-related conditions: An overview of systematic reviews. Phytomedicine 2022, 106, 154430. [Google Scholar] [CrossRef] [PubMed]
  528. Zhu, L.X.; Li, C.Y.; Ji, C.F.; Yang, B.; Li, W.M. The role of substance P and somatostatin in acupuncture and moxibustion-induced postsynaptic inhibition. Zhen Ci Yan Jiu 1993, 18, 290–295. [Google Scholar]
  529. Jin, H.O.; Zhou, L.; Lee, K.Y.; Chang, T.M.; Chey, W.Y. Inhibition of acid secretion by electrical acupuncture is mediated via beta-endorphin and somatostatin. Am. J. Physiol. 1996, 271, G524–G530. [Google Scholar] [CrossRef]
  530. Ruan, H.Z.; Li, X.C.; Li, H.D.; Zhao, B.Y. Somatostatin and electroacupuncture inhibited c-fos expression in spinal cord of arthritic rats. Zhongguo Yao Li Xue Bao 1997, 18, 474–476. [Google Scholar]
  531. Lin, Y.P.; Yi, S.X.; Yan, J.; Chang, X.R. Effect of acupuncture at Foot-Yangming Meridian on gastric mucosal blood flow, gastric motility and brain-gut peptide. World J. Gastroenterol. 2007, 13, 2229–2233. [Google Scholar] [CrossRef] [PubMed]
  532. Zhu, W.L.; Li, Y.; Wei, H.F.; Ren, X.X.; Sun, J.; Zhang, L.F.; Zhu, J. Effect of electro-acupuncture at different acupoints on neuropeptide and somatostatin in rat brain with irritable bowel syndrome. Chin. J. Integr. Med. 2012, 18, 288–292. [Google Scholar] [CrossRef] [PubMed]
  533. Liu, M.R.; Xiao, R.F.; Peng, Z.P.; Zuo, H.N.; Zhu, K.; Wang, S.M. Effect of acupuncture at “Zusanli” (ST 36 and “Taichong” (LR 3) on gastrointestinal hormone levels in rats with diarrhea type irritable bowel syndrome. Zhen Ci Yan Jiu 2012, 37, 363–368. [Google Scholar] [PubMed]
  534. Shao, Y.; Lai, X.S.; Gong, Y.Z.; Yan, B.; He, L.L.; Luo, R.; Tang, C.Z. Effects of electroacupuncture on plasma and cerebral somatostatin and beta-EP contents and learning-memory ability in vascular dementia rats. Zhen Ci Yan Jiu 2008, 33, 98–102. [Google Scholar]
  535. Tian, Q.; Wang, L.; Yao, L.; Zhang, L.; Zhang, H. Effects of Zusanli electroacupuncture on somatostatin expression in the rat brainstem. J. Mol. Neurosci. 2013, 49, 28–37. [Google Scholar] [CrossRef]
  536. Yi, S.X.; Yang, R.D.; Yan, J.; Chang, X.R.; Ling, Y.P. Effect of electro-acupuncture at Foot-Yangming Meridian on somatostatin and expression of somatostatin receptor genes in rabbits with gastric ulcer. World J. Gastroenterol. 2006, 12, 1761–1765. [Google Scholar] [CrossRef]
  537. Zhang, X.; Yuan, Y.; Kuang, P.; Wu, W.; Zhang, F.; Liu, J. Effects of electro-acupuncture on somatostatin and pancreatic polypeptide in ischemic cerebrovascular diseases. J. Tradit. Chin. Med. 1999, 19, 54–58. [Google Scholar]
  538. Wang, P.; Yang, J.; Liu, G.; Chen, H.; Yang, F. Effects of moxibustion at head-points on levels of somatostatin and arginine vasopressin from cerebrospinal fluid in patients with vascular dementia: A randomized controlled trial. Zhong Xi Yi Jie He Xue Bao 2010, 8, 636–640. [Google Scholar] [CrossRef]
  539. Fang, Z. Expression of somatostatin mRNA and coexistence of SOM mRNA and 5-HT in nucleus raphe dorsalis following noxious stimulation and electroacupuncture analgesia. Zhen Ci Yan Jiu 1996, 21, 22–26. [Google Scholar] [PubMed]
  540. Guo, Y.; Luo, R.; Wang, J.; Zhao, Y. Effect of somatostatin on functions of acupuncture meridians. Zhen Ci Yan Jiu 2011, 36, 307–312. [Google Scholar] [PubMed]
  541. Andras, I.; Crisan, D.; Cata, E.; Tamas-Szora, A.; Caraiani, C.; Coman, R.T.; Bungardean, C.; Mirescu, C.; Coman, I.; Crisan, N. MRI-TRUS fusion guided prostate biopsy—Initial experience and assessment of the role of contralateral lobe systematic biopsy. Med. Ultrason. 2019, 21, 37–44. [Google Scholar] [CrossRef]
  542. Cata, E.D.; Andras, I.; Telecan, T.; Tamas-Szora, A.; Coman, R.T.; Stanca, D.V.; Coman, I.; Crisan, N. MRI-targeted prostate biopsy: The next step forward! Med. Pharm. Rep. 2021, 94, 145–157. [Google Scholar] [CrossRef] [PubMed]
  543. Malik, D.; Pant, V.; Sen, I.; Thakral, P.; Das, S.S.; Cb, V. The Role of PET-CT-Guided Metabolic Biopsies in Improving Yield of Inconclusive Anatomical Biopsies: A Review of 5 Years in a Teaching Hospital. Diagnostics 2023, 13, 2221. [Google Scholar] [CrossRef] [PubMed]
  544. Wu, M.-h.; Xiao, L.-f.; Liu, H.-w.; Yang, Z.-q.; Liang, X.-x.; Chen, Y.; Lei, J.; Deng, Z.-m. PET/CT-guided versus CT-guided percutaneous core biopsies in the diagnosis of bone tumors and tumor-like lesions: Which is the better choice? Cancer Imaging 2019, 19, 69. [Google Scholar] [CrossRef]
  545. Cerci, J.J.; Bogoni, M.; Cerci, R.J.; Masukawa, M.; Neto, C.C.; Krauzer, C.; Fanti, S.; Sakamoto, D.G.; Barreiros, R.B.; Nanni, C. PET/CT-guided biopsy of suspected lung lesions requires less rebiopsy than CT-guided biopsy due to inconclusive results. J. Nucl. Med. 2021, 62, 1057–1061. [Google Scholar] [CrossRef] [PubMed]
  546. Popa, G.A.; Preda, E.M.; Scheau, C.; Vilciu, C.; Lupescu, I.G. Updates in MRI characterization of the thymus in myasthenic patients. J. Med. Life 2012, 5, 206–210. [Google Scholar]
  547. Li, H.R.; Gao, J.; Jin, C.; Jiang, J.H.; Ding, J.Y. Comparison between CT and MRI in the Diagnostic Accuracy of Thymic Masses. J. Cancer 2019, 10, 3208–3213. [Google Scholar] [CrossRef]
  548. Hernandez Vargas, S.; Kossatz, S.; Voss, J.; Ghosh, S.C.; Tran Cao, H.S.; Simien, J.; Reiner, T.; Dhingra, S.; Fisher, W.E.; Azhdarinia, A. Specific Targeting of Somatostatin Receptor Subtype-2 for Fluorescence-Guided Surgery. Clin. Cancer Res. 2019, 25, 4332–4342. [Google Scholar] [CrossRef]
  549. Dragosloveanu, S.; Petre, M.A.; Gherghe, M.E.; Nedelea, D.G.; Scheau, C.; Cergan, R. Overall Accuracy of Radiological Digital Planning for Total Hip Arthroplasty in a Specialized Orthopaedics Hospital. J. Clin. Med. 2023, 12, 4503. [Google Scholar] [CrossRef]
  550. Chatalic, K.L.; Kwekkeboom, D.J.; de Jong, M. Radiopeptides for Imaging and Therapy: A Radiant Future. J. Nucl. Med. 2015, 56, 1809–1812. [Google Scholar] [CrossRef]
  551. Fortunati, E.; Bonazzi, N.; Zanoni, L.; Fanti, S.; Ambrosini, V. Molecular imaging Theranostics of Neuroendocrine Tumors. Semin. Nucl. Med. 2023, 53, 539–554. [Google Scholar] [CrossRef]
  552. Acker, G.; Kluge, A.; Lukas, M.; Conti, A.; Pasemann, D.; Meinert, F.; Anh Nguyen, P.T.; Jelgersma, C.; Loebel, F.; Budach, V.; et al. Impact of 68Ga-DOTATOC PET/MRI on robotic radiosurgery treatment planning in meningioma patients: First experiences in a single institution. Neurosurg. Focus 2019, 46, E9. [Google Scholar] [CrossRef]
  553. Dragosloveanu, S.; Petre, M.A.; Capitanu, B.S.; Dragosloveanu, C.D.M.; Cergan, R.; Scheau, C. Initial Learning Curve for Robot-Assisted Total Knee Arthroplasty in a Dedicated Orthopedics Center. J. Clin. Med. 2023, 12, 6950. [Google Scholar] [CrossRef] [PubMed]
  554. Wang, W.; Deng, J.; Li, H.; Ji, Z.; Wen, J. Laparoscopic resection of a paraganglioma behind the retrohepatic segment of the inferior vena cava: A case report and literature review. Front. Endocrinol. 2023, 14, 1171045. [Google Scholar] [CrossRef] [PubMed]
  555. Herrera-Martínez, A.D.; van den Dungen, R.; Dogan-Oruc, F.; van Koetsveld, P.M.; Culler, M.D.; de Herder, W.W.; Luque, R.M.; Feelders, R.A.; Hofland, L.J. Effects of novel somatostatin-dopamine chimeric drugs in 2D and 3D cell culture models of neuroendocrine tumors. Endocr. Relat. Cancer 2019, 26, 585–599. [Google Scholar] [CrossRef] [PubMed]
  556. Timofticiuc, I.A.; Călinescu, O.; Iftime, A.; Dragosloveanu, S.; Caruntu, A.; Scheau, A.E.; Badarau, I.A.; Didilescu, A.C.; Caruntu, C.; Scheau, C. Biomaterials Adapted to Vat Photopolymerization in 3D Printing: Characteristics and Medical Applications. J. Funct. Biomater. 2023, 15, 7. [Google Scholar] [CrossRef]
  557. Periferakis, A.; Periferakis, A.T.; Troumpata, L.; Dragosloveanu, S.; Timofticiuc, I.A.; Georgatos-Garcia, S.; Scheau, A.E.; Periferakis, K.; Caruntu, A.; Badarau, I.A.; et al. Use of Biomaterials in 3D Printing as a Solution to Microbial Infections in Arthroplasty and Osseous Reconstruction. Biomimetics 2024, 9, 154. [Google Scholar] [CrossRef]
  558. Kiseljak-Vassiliades, K.; Xu, M.; Mills, T.S.; Smith, E.E.; Silveira, L.J.; Lillehei, K.O.; Kerr, J.M.; Kleinschmidt-DeMasters, B.K.; Wierman, M.E. Differential somatostatin receptor (SSTR) 1–5 expression and downstream effectors in histologic subtypes of growth hormone pituitary tumors. Mol. Cell. Endocrinol. 2015, 417, 73–83. [Google Scholar] [CrossRef] [PubMed]
  559. Hankus, J.; Tomaszewska, R. Neuroendocrine neoplasms and somatostatin receptor subtypes expression. Nucl. Med. Rev. Cent. East. Eur. 2016, 19, 111–117. [Google Scholar] [CrossRef] [PubMed]
  560. Beaumont, V.; Hepworth, M.B.; Luty, J.S.; Kelly, E.; Henderson, G. Somatostatin receptor desensitization in NG108-15 cells. A consequence of receptor sequestration. J. Biol. Chem. 1998, 273, 33174–33183. [Google Scholar] [CrossRef] [PubMed]
Table 1. Basic properties and distribution of SSTRs.
Table 1. Basic properties and distribution of SSTRs.
ReceptorFirst CloningMR *LengthChromosomal LocationMajor Tissue SitesReferences
SSTR1199245,00039114q13Brain/CNS, gastrointestinal tract, pancreas[27,28,29,30,31]
SSTR2199241,30536917q24Brain/CNS, gastrointestinal tract, pancreas, lymph tissue, adrenal glands[27,32,33,34]
SSTR3199246,00041822q13.1Brain/CNS, pancreas, gastrointestinal tract, lymph tissue, adrenal glands[35,36]
SSTR4199345,00038820p11.2Brain/CNS, retina, placenta[32,37,38,39,40]
SSTR5199439,00036416p13.3Brain/CNS, pancreas, gastrointestinal tract, lymph tissue, adrenal glands, aortic smooth muscle cells, Sertoli cells[5,41,42,43,44]
* MR refers to molecular weight/relative molar mass.
Table 2. Physiological functions of SSTR activation.
Table 2. Physiological functions of SSTR activation.
SSTRFunctionsReferences
SSTR1Inhibition of GH, prolactin and calcitonin secretion, (possible) anti-inflammatory and anti-nociceptive, regulation of hippocampal function[5,20,45]
SSTR2Inhibition of gastrin, histamine, growth hormone, adrenocorticotropin, glucagon, insulin, TSH, interferon-γ secretion, modulation of eating and drinking behavior, inhibition of stress responses, antidepressant effects, retinal neuroprotection[20,46,47,48,49,50,51,52]
SSTR3Cell proliferation reduction and apoptosis induction, insulin release inhibition, growth hormone release inhibition[20,53,54,55,56]
SSTR4Learning and memory, locomotor activity increase, possible modulation of behavioral responses[5,20,57,58]
SSTR5Inhibition of growth hormone, adrenocorticotropin, insulin, glucagon-like peptide-1 and amylase secretion, (possible) anti-stress function, (possible) gastric emptying mediation[5,20,59,60]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Periferakis, A.; Tsigas, G.; Periferakis, A.-T.; Tone, C.M.; Hemes, D.A.; Periferakis, K.; Troumpata, L.; Badarau, I.A.; Scheau, C.; Caruntu, A.; et al. Agonists, Antagonists and Receptors of Somatostatin: Pathophysiological and Therapeutical Implications in Neoplasias. Curr. Issues Mol. Biol. 2024, 46, 9721-9759. https://doi.org/10.3390/cimb46090578

AMA Style

Periferakis A, Tsigas G, Periferakis A-T, Tone CM, Hemes DA, Periferakis K, Troumpata L, Badarau IA, Scheau C, Caruntu A, et al. Agonists, Antagonists and Receptors of Somatostatin: Pathophysiological and Therapeutical Implications in Neoplasias. Current Issues in Molecular Biology. 2024; 46(9):9721-9759. https://doi.org/10.3390/cimb46090578

Chicago/Turabian Style

Periferakis, Argyrios, Georgios Tsigas, Aristodemos-Theodoros Periferakis, Carla Mihaela Tone, Daria Alexandra Hemes, Konstantinos Periferakis, Lamprini Troumpata, Ioana Anca Badarau, Cristian Scheau, Ana Caruntu, and et al. 2024. "Agonists, Antagonists and Receptors of Somatostatin: Pathophysiological and Therapeutical Implications in Neoplasias" Current Issues in Molecular Biology 46, no. 9: 9721-9759. https://doi.org/10.3390/cimb46090578

Article Metrics

Back to TopTop