Next Article in Journal
Proteinoid Nanocapsules as Drug Delivery System for Improving Antipsychotic Activity of Risperidone
Next Article in Special Issue
PET Radiotracers for CNS-Adrenergic Receptors: Developments and Perspectives
Previous Article in Journal
Development of a Molecular Snail Xenomonitoring Assay to Detect Schistosoma haematobium and Schistosoma bovis Infections in their Bulinus Snail Hosts
Previous Article in Special Issue
Evaluation of Met-Val-Lys as a Renal Brush Border Enzyme-Cleavable Linker to Reduce Kidney Uptake of 68Ga-Labeled DOTA-Conjugated Peptides and Peptidomimetics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Overview of Radiolabeled Somatostatin Analogs for Cancer Imaging and Therapy

1
UPS, CNRS, SPCMIB (Laboratoire de Synthèse et Physico-Chimie de Molécules d’Intérêt Biologique)—UMR 5068, Université de Toulouse, F-31062 Toulouse, France
2
Groupement d’Intérêt Public ARRONAX, 1 Rue Aronnax, F-44817 Saint Herblain, France
3
CNRS, CRCINA (Centre de Recherche en Cancérologie et Immunologie Nantes—Angers)—UMR 1232, ERL 6001, Inserm, Université de Nantes, F-44000 Nantes, France
4
Comprehensive Cancer Center Eugène Marquis, Rennes, F-35000, France
5
CNRS, ISCR (Institut des Sciences Chimiques de Rennes)—UMR 6226, Univ Rennes, F-35000 Rennes, France
6
INRAE, Institut NUMECAN (Nutrition, Métabolismes et Cancer)—UMR_A 1341, UMR_S 1241, Inserm, Univ Rennes, F-35000 Rennes, France
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(17), 4012; https://doi.org/10.3390/molecules25174012
Submission received: 28 July 2020 / Revised: 28 August 2020 / Accepted: 1 September 2020 / Published: 2 September 2020
(This article belongs to the Special Issue Radiolabeled Compounds for Diagnosis and Treatment of Cancer)

Abstract

:
Identified in 1973, somatostatin (SST) is a cyclic hormone peptide with a short biological half-life. Somatostatin receptors (SSTRs) are widely expressed in the whole body, with five subtypes described. The interaction between SST and its receptors leads to the internalization of the ligand–receptor complex and triggers different cellular signaling pathways. Interestingly, the expression of SSTRs is significantly enhanced in many solid tumors, especially gastro-entero-pancreatic neuroendocrine tumors (GEP-NET). Thus, somatostatin analogs (SSAs) have been developed to improve the stability of the endogenous ligand and so extend its half-life. Radiolabeled analogs have been developed with several radioelements such as indium-111, technetium-99 m, and recently gallium-68, fluorine-18, and copper-64, to visualize the distribution of receptor overexpression in tumors. Internal metabolic radiotherapy is also used as a therapeutic strategy (e.g., using yttrium-90, lutetium-177, and actinium-225). With some radiopharmaceuticals now used in clinical practice, somatostatin analogs developed for imaging and therapy are an example of the concept of personalized medicine with a theranostic approach. Here, we review the development of these analogs, from the well-established and authorized ones to the most recently developed radiotracers, which have better pharmacokinetic properties and demonstrate increased efficacy and safety, as well as the search for new clinical indications.

1. Introduction

Somatostatin (SST), also called somatotropin release inhibiting factor (SRIF), is a cyclic peptide hormone, first isolated in 1968 from an ovine hypothalamus, and actually identified in 1973 [1]. It was originally discovered as a growth hormone inhibitor, but is now known to be involved in the inhibition of numerous metabolic processes relating to neurotransmitters, endocrine secretions (e.g., growth hormone, insulin, glucagon, and gastrin) but also modulating exocrine secretions (e.g., gastric acid and pancreatic enzymes). In the body, its synthesis takes place in the form of an inactive precursor of 116 amino acids (AA), preprosomatostatin, which is then converted by the action of proteases into prosomatostatin (96 AA). Depending on where it is produced in the body, enzymes do not cleave the pro-peptide on the same amino acid motif, resulting in two distinct active forms, SRIF-28 and SRIF-14. Although SRIF-14 is predominant in the central nervous system and SRIF-28 in the digestive tract, the distribution of these two biologically active forms is similar.
In the early 1990s, concomitantly to studies on the binding properties and mechanisms of action of somatostatin, five receptor subtypes were discovered (SSTR1 to SSTR5) [2]. These subtypes belong to the family of receptors coupled to G-proteins, and their length varies from 364 to 418 AA. They all exhibit seven α helices with transmembrane domains and most of the differences between subtypes are found in the extracellular (N-terminal) and intracellular (C-terminal) ends. SSTR-1, -3, -4, and -5 have a single subtype, while two variants exist for SSTR2, called SSTR2A and SSTR2B. SSTR1 to 4 link SRIF-14 and -28 with a very high affinity (in the nanomolar order), whereas SSTR5 shows an affinity 5 to 10 times higher, but for SRIF-28 only.
Somatostatin receptors are widely distributed in healthy tissues, with distinct expression throughout the body (Figure 1). It is quite possible to find several subtypes in the same tissue. Each of the SSTRs is involved in the regulation of the various processes: (i) SSTR1 is involved in the antisecretory effects of growth hormone, prolactin (a peptide hormone involved in lactation, reproduction, growth, and immunity) and calcitonin (regulation of calcemia); (ii) SSTR2 also inhibits the secretion of growth hormone and adrenocorticotropin (hormone that stimulates the adrenal glands), glucagon, insulin, interferon-γ (protein produced by immune cells), and stomach acid; (iii) SSTR5 has the same inhibiting effect on growth hormone, adrenocorticotropin, insulin, and inhibits the secretion of amylase (digestive enzyme constituting saliva and pancreatic juice); (iv) SSTR3 reduces cell proliferation and causes cell apoptosis; (v) the functions of SSTR4 are not yet well defined [3].
The effects of somatostatin are expressed through different signaling pathways [4,5]. After a cascade of reactions, this leads on the one hand to the inhibition of tumor growth (action on the secretion of hormones) and blocking proliferation via the activation of different tyrosine phosphatases (anti-proliferative and pro-apoptotic action), but also to the inhibition of the secretion of growth factors such as growth hormone or IGF-1 having a major role in the inhibition of tumor growth (anti-angiogenic) (Figure 2) [6,7].
Over the past 20 years, our understanding of the phenomena due to the activation of SSTRs has increased thanks to numerous translational and clinical studies, leading to the development of new therapeutic options [3]. The use of SST analogs has demonstrated real effectiveness in the treatment of various pathologies: acromegaly (production of an excess of growth hormone), pancreatitis, complications linked to diabetes and obesity (e.g., retinopathy or nephropathy), action on inflammation and pain in some cases [5,9]. However, SSTRs and SST analogs are mainly known for their presence and role in the detection and treatment of some solid tumors. Tumor cells and peritumoral vessels express receptor subtypes whose density depends on the type of tumors (Table 1) [10,11,12,13]. For those overexpressing SSTRs, such as pituitary adenomas, gastroentero-pancreatic neuroendocrine tumors (GEP-NET), or other cancers (e.g., lymphomas, small cell lung cancers, etc.), targeting with SST analogs becomes possible [14]. Many therapeutic protocols based on these analogs (classic octreotide or with a longer release time (octreotide LAR), Lanreotide, Vapreotide, Pasireotide, etc.) have been the subject of phase II and III clinical trials. The majority of results were generally disappointing and did not provide clear evidence of a significant antitumor effect on solid tumors, probably due to the existence of other pathways of tumor progression [15,16].
For example, regarding liver tumors, such as hepatocellular carcinoma (HCC), in vitro studies clearly demonstrated (i) the lack of SSTRs expression in healthy liver cells; (ii) overexpression in tumors and metastases of HCC, even though their density is less than in neuroendocrine tumors [21,22]. On the other hand, the results show a heterogeneous expression and strong inter-individual differences. In fact, according to studies, HCCs express high levels of SSTR2 [21,23,24] or SSTR5 [13,19], or even SSTR1 [22] or SSTR3 [25]. In general, around 40% of HCCs studied express somatostatin receptors. These differences could be due to the different methodologies used during the measurements, by studying different stages of the disease or even by heterogeneous behaviors of HCC. Further studies have also found a correlation between the density of SSTRs expression, disease aggressiveness [26], and the rate of tumor recurrence after treatment with octreotide LAR [27]. In a study by Nguyen-Khac et al. [23], 41.2% of extrahepatic metastases express SSTR2. Preclinical tests on HCC cell lines have shown an antiproliferative effect of SST analogs [25,28]. In addition, a real decrease in invasion and cell migration of HCC cells after stimulation of SSTR1 by a specific agonist has also been demonstrated [22]. This action has also been confirmed in vivo [29], with the demonstration of a similar effect on metastatic dissemination [23,30]. These initial results paved the way for clinical trials on patients with HCC, but their conclusions are quite contradictory, [31] showing rather positive effects in the advanced stages [32,33] and others quite negative [34,35]. These outcome discrepancies could come from heterogeneity in the choice of patients, but available data are still insufficient to truly conclude on the effectiveness of analogs of SST alone in the control of HCC tumors [6,31,36]. Cholangiocarcinoma, the other main primary liver tumor, might also be a potential target [18,37].
On the other hand, in certain cases, and in particular for neuroendocrine tumors (a category of tumors where SSTRs are the most expressed), a benefit has been proven via two Phase III studies, which have greatly contributed to the fact that SST analogs are now used in clinical routine [38,39].

2. Somatostatin Analogs

Somatostatin has a short half-life in the body (between one and three minutes), because it is rapidly degraded by peptidases found in plasma and tissues [40]. Therefore, the amount present in the bloodstream is extremely low (between 14 and 32.5 pg/mL). This very short half-life has been considered a limiting factor for possible clinical applications, thus many analogs with better metabolic properties (longer half-life between 1.5 h and 12 h) have been rapidly developed [2,5,9]. These are most often hexapeptide or octapeptide molecules which incorporate the biologically active core of native somatostatin (see some examples in Figure 3). Indeed, studies on the structure–activity correlation have shown that the Phe7, Trp8, Lys9, and Thr10 sequence in the form of a β-sheet is necessary for biological activity. The residues Trp8 and Lys9 are essential for this activity, whereas Phe7 and Thr10 may undergo some substitutions. Among somatostatin analogs, there are two main categories: the agonists (substances capable of activating somatostatin receptors) and the antagonists (molecules that interact with somatostatin receptors and block or reduce the physiological effect of an agonist). It is also important to note that somatostatin analogs have different affinities for the different receptor subtypes [2].
The first agonist peptide analog to be approved by the FDA was octreotide (SMS 201-995), marketed under the name Sandostatin®. From a structural point of view, it has a d-Trp and a d-Phe, to stabilize the β-sheet and a disulfide bridge closer to the active core, for a better metabolic stability. Its pharmacodynamics is highly similar to native SST, which has made it widely used in clinical trials for the treatment of GEP (gastro-entero-pancreatic) tumors [41,42]. Next, Lanreotide (BIM 23014, tradename Somatuline®), whose structure is similar to that of octreotide (Phe and Thr having been replaced by Tyr and Val respectively), showed comparable characteristics and is also widely used in the treatment of neuroendocrine tumors [43]. In 2005, another analog, Vapreotide (RC160), was marketed under the name Sanvar®, with properties close to those of the two previous analogs, and is also used for the treatment of esophageal varices. More recently, Pasireotide (SOM-230 or Signifor®) was one of the first analogs to show a strong affinity for most of the somatostatin receptor subtypes (pansomatostatin analog). Marketed by Novartis, it is used for the treatment of Cushing’s disease [44]. Many other analogs have been developed, from “ultra-short” peptides, such as SDZ 222-100 (an adamantine cyclopeptide), to longer ones, such as KE-108 or CH-275 [5]. Regarding antagonist peptide analogs, the wide variety of compounds that the octapeptide model can offer has allowed the discovery of several structures that can block this kind of receptors. The first antagonist that has been described in the literature is CYN-154806, followed by PRL-2970, sst3-ODN-8 or even non-cyclic models such as BIM-23056 and BIM-23627. New non-peptide compounds have also emerged [45]. These agonists and antagonists (selective or not) constitute a very promising field in the chemistry of somatostatin analogs, in particular because of their pharmacological, pharmacokinetic, and physicochemical properties. This type of compound may have a stronger affinity and/or selectivity for certain subtypes of somatostatin receptors than the majority of peptide analogs. They can thus provide additional information on the exact role of each of these subtypes [5,9].

3. Targeting of Somatostatin Receptors with Radiopharmaceuticals

In the field of medicine, much research is focused on finding methods to achieve earlier detection of pathologies to allow treatment at early stages of the disease, to increase the chances of total recovery. For this purpose, nuclear medicine, through the use of radiopharmaceuticals, is a very powerful tool. Its application can have two different aims: imaging, with the visualization of a radioactive element’s distribution in the body, or therapy, with specific irradiation of abnormal cells, thereby reducing damages to nearby healthy tissue. Having a broad range of potential biological targets and desirable pharmacokinetic characteristics—such as high uptake in target tissue and fast blood and non-target tissue clearance—peptides can also be easily chemically modified for incorporation into a radiopharmaceutical, making them a very potent targeting vector for nuclear medicine. Research in that domain has thus gained widespread interest [46,47,48,49]. These compounds can be directly labeled with a radionuclide, such as a halogen radioisotope, but they are generally based on a triple structure involving: (i) a radiometal, the radiation of which allows either the localization (γ and β+ emitters) or the destruction (β, α or Auger electron emitters) of the targeted cells; (ii) a bifunctional chelating agent (BFCA), the dual role of which is not only to bind the radiometal in a very stable manner to minimize its dissociation in vivo, but also to allow its conjugation with targeting moiety (or vector) via a functionalized arm; (iii) a targeting moiety (the peptide analog), which aims to convey this set in a specific way to a well-defined target. To limit the influence of the chelating moiety, a linker (or spacer) is usually inserted between the BFCA and the biomolecule (Figure 4).
The choice of the radiometal is crucial, since it deeply influences the design of the chelating structure [50,51,52,53]. Several criteria govern the choice of radionuclide: (i) the nature of the radiation emitted, depending on the intended application (diagnosis or therapy); (ii) the half-life, which must be long enough to allow effective fixation of the radiotracer on the target cells, but relatively short to avoid irradiation of the organism (neighboring healthy tissues) and more specifically non-targeted organs; (iii) the isotope decay profile. By emitting its radiation, the nuclide disintegrates into a daughter nuclide, which must be non-radioactive to avoid any additional harmfulness to the organism; (iv) the means of production. Most of the radioelements used in nuclear medicine are artificial. They can be produced in three different ways: from a nuclear reactor, a cyclotron or via a generator. Generator production remains the most convenient way for clinical application, as it can provide in-house radionuclides when a cyclotron is not available nearby, but cyclotron production still remains the cheapest and most used. As an example, Table 2 shows some of the characteristics of radioactive nuclides among the most used today for the radiolabeling of peptides.
From a structural point of view, each radiometal has its own properties such as polarizability, degree of oxidation, or coordination number. These features have a direct impact on the choice of the bifunctional chelating agent, in particular in terms of denticity and nature of the donor atoms (most often O-, N-, or S-donors) [54,55]. The BFCA makes it possible to link the biomolecule and the radiometal; its choice is a crucial step in the construction of a radiopharmaceutical. As indicated above, this structure plays a double role: the first is to complex the radioelement in a very stable manner. Several criteria can be evaluated to truly attest to the stability of the complex formed. First of all, the formed radiocomplex must be thermodynamically stable, i.e., the metal-ligand affinity must be as strong as possible. Then it must be kinetically inert. Many metalation reactions take place in the body and the complex formed must be stable enough to avoid any in vivo degradation (e.g., demetallation or transchelation). In addition, radiolabeling conditions with low concentrations are required, ideally with efficient complexation kinetics (high labeling yield) and fast and mild reaction conditions. Beside chemistry considerations, the radiotracer must have: (i) a strong affinity for the target receptor; (ii) a high accumulation for the target and low for the non-target organs; (iii) relatively rapid clearance in the organism; (iv) preferably a mainly renal route of excretion.
Chelating ligands used for the design of radiotracers are usually classified into two categories: macrocyclic and acyclic compounds (Figure 5). Generally, acyclic ligands are less kinetically inert than macrocycles, although some may have shown very good characteristics. On the other hand, these ligands generally have faster metal-chelate binding kinetics compared to macrocyclic analogs, which represents a huge advantage for working with isotopes that have a short lifespan. Despite the coordination properties specific to each metal, some chelating agents—such as polyaminopolycarboxylic acids—are considered to be ‘universal’ because they can complex different radiometals.
Among acyclic ligands, the first BFCAs developed were EDTA (ethylenediaminetetraacetic acid) and DTPA (diethylenetriaminepentaacetic acid). They have been widely used in the chemistry of radiopharmaceuticals, in particular with radioelements such as 111In, 90Y or 177Lu, and even 99mTc [54]. Later on, DTPA derivatives such as CHX-A′′-DTPA with a cyclohexyl moiety bringing more rigidity to the DTPA backbone (allowing a pre-organization of the system) showed better kinetic inertia [56]. Regarding cyclic compounds, cyclen derivatives such as DOTA (1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetic acid) and triaza analogs—NOTA (1,4,7-triazacyclononane-1,4,7-triacetic acid)—are among the most studied ligands. NOTA has the smaller chelating cavity of the two, and is generally used for Ga (III) or Cu (II) because it has a particular attraction for these metals, which results in mild radiolabeling conditions and good in vivo stability of the complexes formed. DOTA (which is considered as the gold standard chelator) and its derivatives play an important role in clinical applications because they form very stable complexes with a wide range of trivalent radiometals such as Ga (III), Y (III), In (III), Lu (III), or even divalent such as Cu (II) [57,58]. For DOTA or NOTA, the introduction of a functionalized arm offers the possibility of coupling a biomolecule (NODASA/NODAGA and DOTASA/DOTAGA). Similarly, TETA (1,4,8,11-tetraazacyclotetradecane-1,4,8,11-tetraacetic acid), has mainly been studied with Cu (II) and have shown a stability similar to DOTA [59].
Whether on the side of macrocyclic ligands, derivatives or variations of DOTA (e.g., p-SCN-Bn-DOTA, DOTAGA, CB-DO2A, TCMC…), NOTA (e.g., p-SCN-Bn-NOTA, NETA…), or TETA (e.g., CB-TE2A, p-NH2-Bn-TE3A…), or on the side of acyclic ligands, derivatives or variations of DTPA (e.g., CHX-A′′-DTPA…), a large number of ligands have been developed so far. A wide choice of ligands is available for the design of new agents, and numerous journals have described and carefully classified all the structures that can be used in the design of a radiopharmaceutical, whatever the intended application [46,51,53,54,57,60].
BFCA’s second role is to allow the conjugation of the complex with a biomolecule. The nature of this link is very important, because it is essential for it to be stable, and above all, for it to not interfere in any way with the binding to the receiver. The slightest structural modification of the ligand and/or of the biomolecule can have a very marked effect on the affinity to the targeted receptors. To minimize this impact as much as possible, that sometimes a ‘spacer’ or ‘linker’ can be used between these two entities. Biomolecules are often functionalized through a primary amine, which provides an ideal conjugation site for a coupling reaction, most often with peptide or thiourea type links. Other links based on thioether, triazole, oxime, or more recently via a copper-free click-chemistry with tetrazine/cyclooctyne may prove to be interesting, in particular, because they have very good stability in vivo [51,54,61].
Many somatostatin analogs have already been labeled with various radioelements, whether for imaging, with probes used today in clinical applications, or for therapy, with many compounds in clinical studies [17,61,62,63]. These analogs were obtained from modifications in the sequence of amino acids that make up the peptide. For example, replacing Phe3 in octreotide (OC) with Tyr3 (TOC) improves the affinity for SSTRs (in particular SSTR2) and introduction of a Thr (TATE) instead of Thr(ol) (TOC) further improves this. By following this procedure, many analogs have been developed and studied, often with the same chelating cavity to be able to compare their properties (Table 3) [64,65].

3.1. Radiolabeled Somatostatin Analogs for Imaging

The very first proof of concept for the visualization of tumors expressing SSTRs was carried out with [123I-Tyr3]-octreotide, obtained from an iodination reaction (electrophilic substitution) of tyrosine [66,67]. This compound demonstrated biological activity and an affinity for receptors similar to those of native SST [68]. Despite the obvious interest of this probe, several factors such as the difficult radiolabeling procedure, the significant cost, and particularly, the clearance via the liver and the hepatobiliary system (which makes it difficult to interpret the obtained images) were the main drawbacks of its application [67]. To overcome all of these disadvantages, iodine-123 has been replaced with indium-111, which, through the chelating agent DTPA, has been coupled to octreotide (Figure 6) [69]. In vivo studies of [111In-DTPA0]-octreotide ([111In]-pentetreotide) have shown that it is possible to visualize tumors expressing SSTRs and their metastases, even 24 h after injection. In comparison with the compounds coupled to antibodies, this reveals a relatively rapid clearance via the kidneys, which represents a huge advantage compared to [123I-Tyr3]-octreotide [70,71]. This compound was the first radiopharmaceutical targeting SSTRs to be approved by the FDA (Octreoscan® marketed in 1994). It has been widely used, and has long been considered a ‘gold standard’ for the visualization of neuroendocrine tumors. It still has a few limits: in fact, it requires a high tumor/noise intensity ratio, shows low spatial resolution, has a moderate affinity for receptors and finally, and possesses a high γ energy which results in a high dose of radioactivity received by the patient. For all these reasons, research in the field of radiopharmaceuticals has focused on other radioelements such as technetium-99m for SPECT and gallium-68 for PET. In addition to having excellent physical properties, these two elements are available from a commercial clinical-grade generator, an important advantage for clinical applications.

3.1.1. Gallium-68 and Indium-111

DOTATOC analog was the first to be radiolabeled with indium-111, and its comparative study with Octreoscan® showed similar diagnostic accuracy, but with better biodistribution and clearance [72]. Although DOTATATE alone showed better affinity for SSTRs, the two analogs [111In]-DOTATOC and [111In]-DOTATATE showed relatively similar pharmacokinetic properties [73]. SSTR2 receptors—and to a lesser extent, SSTR5—are most often overexpressed in tumors. Consequently, the majority of the radiotracers described have a strong affinity for these two SSTRs subtypes. Systems such as DOTANOC were designed to develop a probe capable of targeting all subtypes. Compared to DOTATOC and DOTATATE, it has a similar affinity for SSTR2 and SSTR5 subtypes, but a much higher affinity towards SSTR3. Their high internalization rate results in interesting biodistribution data, with a greater accumulation of the probe in the tumor and in the organs or tissues expressing SSTRs (e.g., pancreas and adrenal glands), ending with excretion mainly by kidneys [74].
These three systems, similarly labeled with gallium-68 (Figure 7), have proven to be very good radiotracers, and are currently routinely used in clinical applications [75]. These three radiopharmaceuticals have slightly different pharmacokinetic properties, but their diversity is mainly due to the variation in affinity for certain subtypes. This feature is even more marked depending on the radioelement chosen (68Ga or 111In). This can be explained by the differences in the geometry of the complexes. [68Ga]-DOTATOC is very affine for SSTR2 and more moderate for SSTR5, [68Ga]-DOTATATE is specific to SSTR2 and finally, [68Ga]-DOTANOC binds with great affinity to SSTR2, SSTR3, and SSTR5 [76,77,78].
A study with DOTANOC aimed at determining the impact of the introduction of a spacer on the pharmacokinetic properties of the formed radiotracer. The aim was to insert polyethyleneglycol (PEG) moieties or sugars between the chelating cavity (DOTA) and the biomolecule (NOC), which resulted in the modification of the lipophilicity or the charge of the final compound. As a result, the hydrophilicity of the system seems to be involved only in the affinity phenomenon towards the receptor, and the overall charge of the compound influences the excretion profile [79].
DOTA is not the only macrocycle to have been coupled to somatostatin analogs. Knowing the attraction of Ga (III) for NOTA, the latter has been the subject of comparative studies. Conjugated with octreotide (NODAGATOC), the compound showed a strong affinity for SSTR2 (similar to that of DOTATOC). Once marked with 111In, affinity was even stronger for SSTR2, with even a gain on SSTR3 and SSTR5 (compared to 68Ga-NODAGATOC), which confirms the influence that the geometry of the complex can have on affinity. In terms of stability, as expected, that of [68Ga]-NODAGATOC was higher than that of [111In]-NODAGATOC. The biodistribution of [68Ga]-NODAGATOC was similar to that of [68Ga]-DOTATOC, but showed a better accumulation in the tumor than [111In]-DOTATOC. This is probably due to the strong agonist character, and the high rate of internalization of the NODAGATOC derivative [80].
A large variety of derivatives have also been investigated, such as DOTALAN, DOTABOC, DOTAGA [81], DOTANOCATE or DOTABOCATE (all derivatives of DOTANOC) [82,83], or THP-TATE (comparison of the overall behavior of the tris chelating system (hydroxypyridinone) with DOTATATE) [84]. New generation analogs with broader affinity profiles or pan-somatostatin analogs have been developed. For instance, AM3 (DOTA-Tyr-cyclo(DAB-Arg-cyclo(Cys-Phe-d-Trp-Lys-Thr-Cys))), a bicyclic somatostatin analog demonstrated affinity to SSTR2, 3, and 5, when labeled with 68Ga. It showed a fast background clearance coupled with a high tumor/non-tumor ratio. [85] KE108 was coupled with DOTA and labeled with 111In and 68Ga, giving [111In/68Ga]-KE88 (DOTA-d-Dab-Arg-Phe-Phe-d-Trp-Lys-Thr-Phe), which bound to all five SSTRs with high affinity. [86] However, in an in vitro study, it had a low SSTR2 uptake, but was very effective for SSTR3-expressing tumors. More recently, a Pasireotide derivative, DOTA-PA1 (DOTA-cyclo-[HyPro-Phe-d-Trp-Lys-Tyr(Bzl)-Phe]) was labeled with 68Ga and was investigated in three human lung cancer models, where it demonstrated superiority compared to [68Ga]-DOTATATE [87]. In parallel, the group from Demokritos Institute, in Athens, developed pansomatostatin radiopeptides based on native somatostatin (SRIF-14 and SRIF-28). Both were derivatized with DOTA chelator and labeled with 111In. Subsequent radiotracers exhibited high affinity and internalization profiles. SRIF-14 derivatives unfortunately demonstrated low in vivo stability. [111In]-DOTA-LTT-SS28, on the contrary, demonstrated a much higher stability and showed more promise [88,89].

3.1.2. Technetium-99m

A wide range of chelating agents have been used to prepare somatostatin analogs labeled with technetium-99m: peptide moieties [90,91], propyleneaminooxime [92], tetraamines [93,94] or a cyclopentadienyl group [95]. Macrocyclic ligands have also been investigated [96]. Three systems stand out for the radiolabeling of somatostatin analogs: HYNIC-TOC and Demotate scaffolds, and P829 (Figure 8).
Initially, the HYNIC core (hydrazinonicotinamide) was designed for the radiolabeling of antibodies and proteins with technetium-99m [97], then this was transposed to peptides and more specifically to octreotide. This ligand can complex the metal in a monodentate or bidentate way, therefore, it is necessary to use one or more co-ligands to complete the coordination of the [99mTc]-HYNIC core. Among the most commonly used co-ligands are tricin, nicotinic acid, or EDDA (ethylenediaminodiacetic acid). Each co-ligand has its own influences on the properties of the complex obtained (e.g., lipophilicity and biodistribution) [98]. The first studies were carried out using tricin as a co-ligand ([99mTc]-HYNIC-TOC), but quickly EDDA demonstrated a very favorable influence on the pharmacokinetics of the radiotracers [99]. Compared to Octreoscan®, [99mTc]-EDDA/HYNIC-TOC showed better accumulation in the tumor and a weaker accumulation in the kidneys. The improved spatial resolution, the reduction in the radiation dose received by the patient and the better availability of 99mTc made it a possible alternative to Octreoscan® [99,100]. Finally, its conjugation with the octreotate analog ([99mTc]-EDDA/HYNIC-TATE) has shown significantly similar behavior to its octreotide counterpart [101]. [99mTc]-EDDA/HYNIC-TOC (Tektrotyd®) was granted marketing authorization in Europe in adult patients with gastro-enteropancreatic neuroendocrine tumors (GEP-NET) for localizing primary tumors and their metastases.
The second radiotracer, based on the tetraamine motif 6-R-1,4,8,11-tetraazaundecane, is available in a series with [99mTc]-Demotate 1 ([99mTc-N40, Tyr3]-octreotate) and 99mTc-Demotate 2 ([99mTc-N40–1, Asp0, Tyr3]-octreotate). The first version of this probe demonstrated excellent pharmacokinetic properties, including faster accumulation in the tumor compared to Octreoscan® [102]. The objective of the second version was to improve the qualities of [99mTc]-Demotate 1, by modifying the overall charge of the complex and adding an Asp residue. In the end, [99mTc]-Demotate 2 showed overall behavior similar to [111In]-DOTATATE, even if the latter has a faster clearance and a better retention time in the tumor [103]. The last of the main analogs based on technetium-99m is [99mTc]-P829 (99mTc-Depreotide), marketed in 2000 by the company CISBio International under the name of NeoSpect®, but recently withdrawn from the market. The P829 peptide (directly radiolabeled with 99mTc) showed results similar to the other SST analogs [104]. Its use for the detection of neuroendocrine tumors appeared to be less precise than with Octreoscan® [105]. On the other hand, its affinity for SSTR3, subtype which may be the origin of cross-competition from other types of receptors (notably VIP receptors), gave it the ability to bind to a larger number of primary tumors [104]. In particular, it was used clinically for the diagnosis of malignant lung tumors [106,107,108], for which it got its market authorization [109], and also demonstrated some interest in breast cancer, but it was never confirmed in a larger series of patients [110].
The question that now remains to be answered is that of the clinical interest of a SPECT tracer among the wide choice of PET SSTRs imaging agents [111,112].

3.1.3. Copper-64

Due to the short half-life of 68Ga (T1/2 = 67.7 min.) each center willing to perform 68Ga PET imaging must purchase a currently expensive 68Ge/68Ga generator and a specifically shielded hot-cell. For this reason and despite the FDA and EMA market authorizations for [68Ga]-DOTATATE and [68Ga]-DOTATOC and the better diagnostic performances for these two radiopharmaceuticals products, the use of 68Ga appears to be under the dependence of an economic choice for many hospitals and only a few large centers are making the financial investment to perform 68Ga-radiolabeling. In this context, the use of a PET-emitter with a longer half-life such as copper-64 (T1/2 = 12.7 h) appears to be an interesting alternative to remove the financial hindrance of gallium-68 [113]. This physical parameter allows for a centralized radiolabeling site with a large multicentric supply of ready-to-use 64Cu-radiolabeled compounds. The chemistry of copper is also well known, which is a real asset in the design of new radiotracers. Many systems already presented before, such as DOTATOC/TATE or NODAGATOC/TATE, or others more copper-specific BFCAs, such as TETA (1,4,8,11-tetraazacyclotetradecane-N,N′,N″,N‴-tetraacetic acid) [114], and its more stable derivatives such as cross-bridge CB-TE2A (4,11-bis(carboxymethyl)-1,4,8,11-tetraazabicyclo [6.6.2]hexadecane) [115], and CPTA (4-[(1,4,8,11-tetraazacyclotetradec-1-yl)methyl]benzoic acid]) [116] or sarcophagine derivatives [117] have been studied. A review on the development of copper radiolabeled somatostatin analogs was recently published by Marciniak et al. [118].
To validate the clinical interest of [64Cu]-somatostatin analogs, various clinical studies have been conducted around the world. Among the different somatostatin analogs, [64Cu]-DOTATATE was one of the first used. In 2015, [64Cu]-DOTATATE was compared head-to-head to [111In]-DTPA-octreotide in 112 patients and showed that the PET 64Cu-compound was far superior to SPECT 111In compound performances [119]. In 2017, [64Cu]-DOTATATE was challenged to [68Ga]-DOTATOC according to an identical PET/CT imaging modality [120]. The results of this study, where 59 patients were injected with [68Ga]-DOTATOC followed by an injection of [64Cu]-DOTATATE one week later, concluded that the two radiopharmaceuticals had the same sensitivity. Nevertheless, in this cohort of neuroendocrine tumors, [64Cu]-DOTATATE had a substantially better lesion detection rate. The patient follow-up revealed that these additional lesions detected by [64Cu]-DOTATATE were true positives. To evaluate the benefits of this better detection of lesions with [64Cu]-DOTATATE than with [68Ga]-DOTATOC, the correlation between PET image [64Cu]-DOTATATE uptake (expressed in maximal standardized uptake value - SUVmax) and overall (OS)/progression free survival (PFS) was studied during 24 months after [64Cu]-DOTATATE PET/CT acquisition. The conclusion of this study claimed a good correlation/prognostic between SUVmax and PFS but not with OS [121]. The major drawback of these preliminary human studies consist of the affinities differences for the five SSTRs subtypes between DOTATOC and DOTATATE compounds. To circumvent these discrepancies, an in vitro study in a mouse model was conducted and compared [64Cu]-DOTATATE to [68Ga]-DOTATATE. The results showed a similar pharmacokinetic and absolute uptake between both compounds 1 h post-injection [122]. In Europe, where the PET radiopharmaceutical approved is [68Ga]-DOTATOC, it could be interesting to perform some PET imaging with [64Cu]-DOTATOC to compare the performance of the two tracers. A first-in-human retrospective study was recently conducted and seems to present same results than [64Cu]-DOTATATE with high detection rate of suspected lesion associated to a high target-to-background contrast [123]. A recent first-in-human study also demonstrated potential interest for [64Cu]-SARTATE analog [124].
In conclusion, despite a higher dosimetric impact for copper-64 (only 17.6% of radioactive decay lead to positron emission), copper-64 somatostatin analogs appear to be an advantageous alternative to gallium-68 radiopharmaceuticals. Compared to 68Ga, in addition to economic advantages, 64Cu has a lower positron range which leads to a better PET intrinsic resolution and a higher half-life which allows for a more flexible scanning window. The better patient care management and outcomes remain to be proven and the work is in progress to establish these points [121,125]. In parallel, at present, a radiopharmaceutical industrial company submitted a market authorization from FDA for [64Cu]-DOTATATE and thus confirms the interest of copper-64 in SSTRs imaging.

3.1.4. Other Radiometals

Other radionuclides have also been investigated for SSTRs imaging. Cobalt-55 seems to be a possible alternative to gallium-68 and copper-64 compounds, with similar behavior and lifespan (17.5 h vs. 12.7 h) to the latter, but with a higher positron yield (75.9% vs. 17.6%). Preliminary complexation tests of DOTATOC with the isotope 57Co as a surrogate for 55Co showed a higher affinity for SSTR2 than [68Ga]-DOTATOC, implying a rate of internalization among the highest of all derivatives of SST and thus, a strong accumulation in targeted tissues. Despite similar structures, the analogs of cobalt and gallium have different biological behaviors. This confirms the fact that the physical characteristics of radioactive elements influence the affinity, biodistribution, and pharmacokinetics of radiolabeled peptides [126]. The properties of cobalt-based compounds have been further investigated with the comprehensive evaluation of other octreotide analogs such as DOTANOC and DOTATATE [127]. Furthermore, [55Co]-DOTATATE compared favorably with [68Ga]-DOTATATE and [64Cu]-DOTATATE in an animal model [122]. Associated with the Auger-emitting 58mCo, it could represent a potentially interesting theranostic pair [128].
Scandium and terbium are two metals that recently emerged as possibly useful for theranostic applications, as both possess imaging and therapeutic radionuclides [129]. DOTATOC was radiolabeled with scandium-44 (T1/2 = 3.97 h, Eβ+ = 632 keV) [130] and terbium-152 (T1/2 = 17.5 h, Eβ+ = 1140 keV) [131] and rapidly injected in patients in proof-of-concept studies [132,133]. No adverse effects were observed during follow-up periods and images proved suitable for diagnosis. With DOTATATE, it seems the affinity to SSTR2 receptors is lower with scandium than with gallium, thus limiting its interest [134]. In a study comparing the labeling and stability of DOTANOC and NODAGANOC with 44Sc and 68Ga, it was observed that [44Sc]-NODAGANOC labeling was more challenging and less stable than [44Sc]-DOTANOC [135]. The opposite was observed with 68Ga. Recently, a new chelator was proposed, AAZTA (1,4-bis (carboxymethyl)-6-[bis (carboxymethyl)]amino-6-methylperhydro-1,4-diazepine), which enables fast and easy labeling at room temperature. AAZTA-TOC labeled with 44Sc demonstrated high in vitro stability [136]. Affinity tests are now necessary to assess its potential utility. DOTATATE has also been labeled with 155Tb (T1/2 = 5.32 days, Eγ = 87 keV (32%), 105 keV (25%)) for SPECT imaging [137]. Though a potentially promising radionuclide for theranostic applications, availability of 155Tb is currently the main limitation for further development.
At the turn of the millennium, yttrium-86 (T1/2 = 14.74 h, 32% β+) was thought to be a potential radionuclide of interest, particularly for pretherapeutic dosimetry of 90Y-radiotracers, and notably 90Y-labeled somatostatin analogs [138]. Thus, several octreotide analogs were developed [139,140]. [86Y]-DOTATOC even reached the clinics [139,141]; however, 86Y properties are less than optimal, and availability is limited, so interest soon faded out.

3.1.5. Fluorine-18

Radiometals’ production is currently still limited, even for the most advanced ones [142,143,144]. Fluorine-18, on the contrary, can be mass-produced and distributed daily, thanks to a worldwide network of cyclotrons. Because of this availability, and favorable decay characteristics (T1/2 = 110 min, 97% β+), it thus should be noted that some radiotracers based on fluorine-18 have been described (Figure 9) [145]. The first generations such as 2-[18F]fluoropropionyl-d-Phe1-octreotide [146] or 4-[18F]fluorobenzoyl-d-Phe1-octreotide [147] generally showed unfavorable biokinetic properties (low accumulation and low retention in the tumor). The probes developed subsequently contained hydrophilic or charged moieties to reduce the lipophilicity of the radiotracer. In particular, several carbohydrate derivatives of octreotide/octreotate have been developed [148,149]. A disadvantage of fluorine-labeling compared to radiometal labeling is the use of generally long and tedious multi-step procedures. To circumvent this, innovative strategies, enabling fast and purification-less labeling, have been developed, such as the formation of 18F-boron or 18F-silicon bonds, or the use of click-chemistry [150,151,152]. Another elegant method to label somatostatin analogs is the use of [18F]-aluminum fluoride with radiotracers previously developed for radiometals, such as NOTATOC [153]. These new generation analogs demonstrated general properties (affinity for the targeted receptors, metabolic stability, biodistribution and clearance) which are much more interesting, and some of them have been investigated in patients, where they gave results comparable to [68Ga]-DOTATOC [154,155]. In addition, [18F]F-FET-βAG-TOCA and [18F]-IMP466 ([Al18F]-NOTATOC) are currently being evaluated in phase I clinical trials (EudraCT number 2013-003152-20 and NCT03511768, respectively). Recently published results with [18F]-IMP466 demonstrated it was safe and well-tolerated, with a physiologic uptake pattern similar to [68Ga]-DOTATATE [156]. Besides cost and availability, another advantage of fluorine-18 is its shorter positron range compared with gallium-68, leading to an improved spatial resolution, and thus, better quantification of uptake [157].

3.2. Radiolabeled Somatostatin Analogs for Therapy

Concerning radionuclide therapy and more particularly peptide receptor radionuclide therapy (PRRT), radioactivity is used to destroy the targeted cells. Radiopharmaceuticals used in therapy are designed in the same way as those used in imaging, only the nature of the radioelement being modified. Contrary to imaging, which uses radioelements having very penetrating but little ionizing radiations, PRRT privileges the use of radionuclides that have little penetrating and more energetic and thus more ionizing radiations. Brought directly to the cancer cell, the radiation emitted by the radioactive decay causes irreversible ionization of the cell’s DNA, which induces its apoptosis. The main isotopes used today are iodine-131, yttrium-90, lutetium-177 and, to a lesser extent, rhenium-188 [158]. As mentioned earlier, the purpose of the DOTA-SSA design was to work with a chelating cavity capable of complexing radioelements for imaging or therapy. Consequently, most of the platforms discussed above have been transposed for therapeutic application via the use of β emitters [64,74,81,82].

3.2.1. Yttrium-90 and Lutetium-177

Yttrium-90, a pure high energy β emitter (T1/2 = 64 h, Eβmax = 2.28 MeV), and lutetium-177, a medium energy β emitter (T1/2 = 6.7 d, Eβmax = 0.5 MeV) with a γ component (208 keV), are currently the most used in PRRT. Each of these two elements has its own advantages for targeted therapy. The particles emitted by 90Y are more energetic and more penetrating; they are able to diffuse on a thicker layer of cells, which is an advantage for the treatment of large tumors. However, even if high energy radiation allows a more uniform irradiation of the tumor, the risk of imposing an excessive dose of radiation on the adjacent tissues is very present. For its part, the 177Lu emits less energetic radiation, more suited to small tumors. In addition, the energy of its γ radiation is sufficient to allow detection by scintigraphy and establish dosimetry during the therapy sequences [159].
The first analog to be studied was [90Y]-DOTATOC (Octreother®), and the first treatment sessions quickly showed good results, stopping the progression of the tumor [72,160,161]. Many studies on this long-used treatment have made it possible to observe a good tolerance for this radiotracer, with fairly mild side effects (fatigue) and in very rare cases a little more severe ones (nausea). However, it also showed some toxicity for the kidneys and the bones, these two aspects being the dose-limiting factors for the patient. In vitro, a greater affinity for SSTR2 has been demonstrated for [90Y]-DOTATATE compared to [90Y]-DOTATOC [64]. However, for the diagnosis in humans, a better contrast between the kidneys and the tumor was found for [111In]-DOTATOC compared to [111In]-DOTATATE [73], which may explain the wider use of DOTATOC analog. Despite this, these two analogs have relatively similar properties and have proven to be effective treatment methods that improve survival in some patients with neuroendocrine tumors (approximately 50 months vs. 18 months without treatment) [162]. In a Phase IIA study with [90Y]-DOTALAN (MAURITIUS trial), this one demonstrated lower tumor uptake in neuroendocrine tumors compared to 90Y-DOTATOC, but could be of potential interest for other tumors, such as HCC or lung cancers [163]. With the perspective of several years of clinical use, PRRT with 90Y-labeled somatostatin analogs appears to be well-tolerated with favorable long-term outcome. Unfortunately, Phase III studies are still lacking [164,165].
The same analogs have also been radiolabeled with lutetium-177. Initially, [177Lu]-DOTATOC was used in cases of relapse of neuroendocrine tumors after treatment with [90Y]-DOTATOC. Despite satisfactory results [166], its subsequently developed analog [177Lu]-DOTATATE has shown more promise, mainly due to a more significant retention time in the tumor. For this reason, octreotate analog (TATE) is being preferred to octreotide (TOC) for labeling with lutetium [164,167]. It is also important to note that, unlike 90Y, no cases of nephrotoxicity after treatment with 177Lu have been reported. In 2005, the possibility of combining these two β emitters for therapy in cases where tumors of variable sizes are detected, was demonstrated [168]. From there, different treatment combinations between the four main systems ([90Y]-DOTATOC, [90Y]-DOTATATE, [177Lu]-DOTATOC, and [177Lu]-DOTATATE) have proven to be interesting and sometimes even more effective than using a single treatment modality [169,170]. Similarly, combination treatments with non-labeled somatostatin analogs, chemotherapy, targeted therapy, and/or radiosensitizers might further improve the efficacy and/or tolerability [171,172]. [177Lu]-DOTATATE has been investigated in a phase III trial, in well-differentiated, unresectable or metastatic, progressive midgut neuroendocrine tumors (Netter 1 trial). Treatment with [177Lu]-DOTATATE resulted in a significant tumor response rate of 18% compared with 3% in the high-dose octreotide LAR group, coupled with a 79% risk reduction for disease progression or death [173]. Following these positive findings, [177Lu]-DOTATATE was granted marketing authorization in this indication, both in Europe and in the US (Lutathera®) [174]. Coupled with 68Ga-imaging (Figure 10), it represents a powerful theranostic tool for the management of neuroendocrine tumors (NETs) [175]. Current research with [177Lu]-DOTATATE aims to improve the safety and efficacy of this procedure, enlarge possible indication, notably in advanced, poorly-differentiated, GEP-NETs, [176,177] or other NETs, such as pheomochromocytoma or paraganglioma [178,179].

3.2.2. Rhenium-188 and Other β-Emitting Radionuclides

Despite equally interesting characteristics, rhenium-188 remains widely less used than 90Y and 177Lu [180]. This is mainly due to more difficult chemistry and the unavailability of a pharmaceutical-grade 188W/188Re generator, as compared to the other two. Vapreotide and Lanreotide analogs have been described in the literature with 188Re. They have been investigated in experimental cancer models (e.g., pancreas, colorectal, lungs and cervical) to reduce tumor growth [181,182,183,184]. [188Re]-Lanreotide notably demonstrated favorable pharmacokinetics and distribution profiles (tumor-to-liver ratio) in HCC-bearing rats compared to healthy ones [185]. Another example is an equivalent to Depreotide (P829). After the development of 99mTc-Depreotide for imaging, the idea was to label this compound with 188Re, to assess its potential in vivo. Although the radiolabeling proceeded successfully, the study showed unacceptable toxicity to non-target organs. To improve its properties, structural modifications of the peptide sequences close to the chelating moiety were tested. This optimization led to P2045, which showed better accumulation in the tumor, weaker retention in the kidneys, and faster urinary excretion than [99mTc]-depreotide [186]. This new rhenium-based analog of depreotide, [188Re]-P2045 (Figure 11), went up to phase I in therapy for small cell lung cancer [187] and has shown promising in vivo results in the treatment of pancreatic tumors in mice [188]. To the best of our knowledge, no HYNIC-TOC/TATE or demotate derivatives have yet been radiolabeled with rhenium. Recent research with rhenium isotopes has been focusing on tricarbonyl core derivatives for the labeling of NOTA-SSAs [96].
In a theranostic perspective, other β-emitting nuclides could have a potential interest—such as 47Sc (T1/2 = 3.35 d, Eβmax = 600.8 keV), 67Cu (T1/2 = 2.58 d, Eβmax = 577 keV), and 161Tb (T1/2 = 6.91 d, Eβmax = 593 keV)—to be coupled with 44Sc, 64Cu, and 152Tb/155Tb respectively [129,158,189]. To date, no 67Cu-labeled somatostatin analogs have been described so far, and only very preliminary studies have been described with [161Tb]-DTPA-Octreotide and [47Sc]-DOTATOC [190,191].

3.2.3. Alpha and Auger Emitters

Recently, alpha emitters have attracted particular attention for radionuclide therapy. Long confined to hematological tumors, they are now being considered for the potential treatment of solid tumors [192]. In vitro, α-labeled somatostatin analogs (DOTATOC and DOTATATE) demonstrated a significantly higher killing effect compared to 177Lu [193,194,195]. [213Bi]- and [225Ac]-labeled DOTATOC (213Bi: T1/2 = 45.6 min, Eα = 5.88 MeV; 225Ac: T1/2 = 9.92 d, Eα = 5.83 MeV) have demonstrated promising therapeutic effects in pre-clinical animal studies [196,197]; whereas [213Bi]-DOTATATE, investigated in human small cell lung carcinoma and rat pancreatic tumor models, demonstrated a great therapeutic effect in both small (50 mm3) and large (200 mm3) tumors, but with a higher probability for stable disease in small tumors [198]. First, and, to date, the only clinical experience with [213Bi]-DOTATOC, was published by Kratochwil et al., and included seven patients with advanced NETs with liver metastases refractory to treatment with [90Y]-DOTATOC or [177Lu]-DOTATOC [199]. It demonstrated specific tumor binding, lower toxicity than with β-irradiation and partial remission of metastases. Two years after intra-arterial injection of [213Bi]-DOTATOC, all seven patients were still alive. Regarding 225Ac, a first-in-human study included 10 patients with progressive NETs after β-PRRT. As with 213Bi, [225Ac]-DOTATOC was well tolerated and effective [200]. A recent study with [225Ac]-DOTATATE confirmed the potential of these radiotracers as an additional, and valuable, treatment option for patients who are refractory to [177Lu]-DOTATATE therapy. 32 patients with previous [177Lu]-DOTATATE therapy were treated with [225Ac]-DOTATATE (100 kBq/kg body weight). The response was assessed in 24 patients, with 9 stabilized diseases and 15 partial remissions [201].
Though not stricto sensu an α-emitter, lead-212 (T1/2 = 10.6 h) eventually decays to stable 208Pb through a cascade chain with two α-emissions of potential therapeutic interest. A somatostatin analog, DOTAMTATE (Figure 12), has been labeled with 212Pb and investigated in a murine model of neuroendocrine tumor. Results showed a promising safety index with a 3.2-fold increase in median survival and one-third of the animals being tumor-free. A combination with 5-FU (Fluorouracyl) was able to durably cure approximately 80% of the animals. [202] Given these promising outcomes, a Phase I dose-escalation clinical trial has recently been started with [212Pb]-DOTAMTATE (AlphaMedix™) including 50 patients with unresectable or metastatic neuroendocrine tumors (NCT03466216). Preliminary results (nine patients enrolled) demonstrated a favorable safety profile at the tested doses [203].
Cyclotron-produced astatine-211 (T1/2 = 7.2 h, Eα = 5.87 MeV) is another very promising α-emitting radionuclide. Astatine is the heaviest halogen with a behavior somehow similar to iodine, but, in certain circumstances, it also displays significant metallic characteristics [204]. Direct astatination of somatostatin analogs is feasible, through tyrosine residues, but it led to poor stability of the resulting analogs, therefore different prosthetic groups have been developed [205,206,207]. Although N-(3-[211At]astato-4-guanidinomethylbenzoyl)-Phe1-octreotate ([211At]-AGMBO) and Nα-(1-deoxy-d-fructosyl)-Nε-(3-[211At]astatobenzoyl)-Lys0-octreotate ([211At]-GABLO) showed disappointing biodistribution results, with poor tumor uptake, [211At]-SPC-octreotide displayed a more favorable biodistribution profile, and a dose-dependent apoptosis in an NSCLC murine model.
Auger electron emitters are also very potent for specific tumor cell killing, sparing surrounding cells, with a highly localized energy deposition. Indium-111 emits Auger electrons (EAe- = 19 keV, 16%), and, as such, has been investigated for therapy. Several clinical trials have been undertaken with high doses of [111In]-Pentetreotide. A first study with 20 patients that had neuroendocrine progressive tumors demonstrated stabilization of the disease in 5 patients, and tumor shrinkage in 5 others. All of them had received a cumulated dose higher than 20 GBq [208]. In a study with 50 SSTR-positive patients treated with cumulated doses from 20 to 160 GBq, of which 40 were evaluable, there was a stabilization in 14 patients, minor remission in 6 and partial remission in 1, with mild bone marrow toxicity [209]. However, half of the patients receiving more than 100 GBq developed a myelodysplastic syndrome or leukemia. A dose of 100 GBq was thus considered the maximal tolerated activity. Another study with 27 patients with GEP-NETs found that two doses of 6.6 GBq (180 mCi) were safe and well-tolerated, demonstrating a clinical benefit in 62% of patients [210]. Benefit of 111In-Pentetreotide treatment was shown to last at least 6 months for 70% of patients, while only 31% of them still had sustained benefit after 18 months [211]. Efficacy in large tumors and end-stage patients is limited, mainly because of heterogeneous radiopharmaceutical uptake due to poor tumor vascularity and central necrosis [212]. This has been demonstrated by Capello et al. in a rat tumor model, with different sizes of tumors [213]. Effects were much more pronounced in small (≤ 1 cm2) tumors than in large (≥8 cm2). They also found a significant increase in tumor receptor density after tumor regrowth, indicating repeated injections would probably be more efficient than single-dose treatment. It could also be worth using PRRT with Auger emitters in an adjuvant setting after surgery, to destroy occult metastases. A final example is [58mCo]-DOTATOC. This radiotracer presented for potential use in Auger-based therapy, particularly for disseminated tumor cells and micrometastases, appears to have more beneficial in vitro properties than those of [177Lu]-DOTATATE, with a significantly more efficient cell killing effect per cumulated decay, which has to be confirmed in vivo [127].

4. Antagonists vs. Agonists

Pharmacomodulation around the synthetic somatostatin analogs has led to a change of chirality in the first amino-acid (from d to l form) and in cysteine number 2 (from l to d form). These modifications have given a new class of SSTR specific compounds with antagonist effects (Table 4). From a pharmacological point of view, the biological and molecular mechanisms responsible for their targeting effectiveness in vivo are completely different. After binding to an SST receptor, an agonist analog is internalized into the cell as a ligand-receptor complex. This internalization allows it to accumulate in the cell, and to increase the amount of radiation emitted. This very powerful and specific internalization mechanism enables efficient in vivo targeting of receptors. This phenomenon does not occur (or very little) for somatostatin antagonists, and they do not stimulate the G-protein coupled to the SSTR with an associate blockage of the agonist-induced activity. Surprisingly, it has been shown that targeting receptors can also be effective without internalization of the ligand-receptor complex, and some antagonist analogs can sometimes behave better than agonists (e.g., better accumulation in tumor, poor kidney retention, and rapid clearance) [214,215]. This high tumor uptake appears to be a consequence of a greater number of target binding sites for antagonists and a more slowly dissociation than for agonists, which allows for a longer accumulation of radiation [216,217]. The hypothesis of a ligand rebinding mechanism has been put forward, but this still requires some investigation before it can be validated. These first results were confirmed by preclinical studies and by preliminary clinical trials and seems to show superior results for antagonist-based tracers than agonists [218,219,220,221]. The first comparative study of antagonists with Octreoscan® confirmed the good characteristics of the [111In]-DOTA-BASS analog, and better accumulation at the level of the tumor and better visualization of metastases. It was truly the first proof of the concept of antagonist SSTRs imaging [222].
Concerning the affinity for each SSTR subtype, it turned out that the nature of the chelator and the radiometal is of great importance for the in vivo pharmacokinetic fate (mainly for the tumor uptake and retention time) [223]. Ultimately, copper-64 based radiotracers seem to be more interesting, especially when comparing their contrast ratio between the tumor and normal tissues which increases over time—a direct consequence of their higher half-life. The influence of radiometals (111In, 90Y, 177Lu, 64Cu, and 68Ga) and chelates (DOTA and NODAGA) on three antagonist families (LM3, JR10, and JR11) were also studied. On the radiometric side, the overall affinity of [68Ga]-DOTA was found to be much lower than for the other elements, which is the opposite of the results obtained with the agonists. For the chelate, the substitution of DOTA by NODAGA seems to greatly improve the affinity of the antagonist analogs. During this study, two particularly promising platforms emerged, DOTA-JR11 and NODAGA-JR11 [224]. Another example highlighting the influence of the chelate is 406-040-15 (cyclo (2–11) H-Cpa-DCys-Asn-Phe-Phe-DTrp-Lys-Thr-Phe-Thr-Cys-2NalNH2), a pansomatostatin analog, with an SSTR3 antagonist behavior. Chelation to DOTA turned this analog to an agonist [225]. Note that the first antagonist labeled via a [99mTc]-tricarbonyl core has been described. 99mTcL-sst2-ANT (with L = tridentate ligand type N, S, N) has shown very promising in vivo behavior, but requires some modifications to improve its pharmacokinetics [226].
As for imaging, antagonists are also an interesting alternative for therapy. As discussed above, the first proof of the feasibility of imaging using antagonists was highlighted by comparing Octreoscan® and [111In]-DOTA-BASS. However, this analog has shown only a very modest affinity for the SSTR2 receptor subtype targeted in the therapy of neuroendocrine tumors [214]. To overcome this problem, the second generation of somatostatin antagonists was synthesized to improve affinity for this receptor. DOTA-JR11 showed the highest affinity for SSTR2 and was selected for use in targeted therapy [218]. A pilot study to assess the possibility of treatment with [177Lu]-DOTA-JR11, by comparing it to [177Lu]-DOTATATE, was carried out. This new antagonist has shown favorable properties, such as better accumulation in the tumor and a higher dose received by the tumor, thanks to a longer retention time [227]. Further developments led to a theranostic pair with JR11: one with a NODAGA chelator (satoreotide trizoxetan, OPS-202) and one with DOTA chelator (satoreotide tetraxetan, OPS-201) [228,229]. Satoreotide trizoxetan is currently radiolabeled with 68Ga and used in PET imaging clinical trials (Figure 13) [230,231]. Satoreotide tetraxetan radiolabeled with 177Lu has been evaluated in a therapeutic clinical trial [232]. First clinical results for this somatostatin antagonist theranostic pair seem to be promising with high sensitivity for neuroendocrine tumors and require further studies in larger patient population.

5. Future Prospects

Regarding clinically established somatostatin analogs, the development of kit-based 68Ga radiotracers, as well as cyclotron production of gallium-68 should improve their availability and worldwide dissemination. Further clinical translation of 64Cu- and 18F-based somatostatin SSAs could also represent an attractive alternative. For therapy, current research focuses on optimizing the dose received by the tumor while sparing healthy tissues. Fractionation, as well as combination of 90Y and 177Lu, have demonstrated their interest [168,234]. The same approach with other treatment modalities, such as external-beam radiotherapy or chemotherapy could enhance treatment response [235,236]. Targeted α-therapy also seems to hold promises and is currently attracting much interest, notably from the industry.
Recent developments showed a switch from agonist to antagonist derivatives, demonstrating higher efficacy. With the advent of new promising radionuclides and somatostatin analogs with better pharmacokinetic properties and binding profiles, the future looks bright for radiolabeled somatostatin analogs, expanding their use for wider indications, than just GEP-NETs. With peptide derivatives with improved targeting, tumors with lower SSTR expression might nonetheless be clinically relevant. In this context, as already demonstrated with some analogs, use of somatostatin-based radiopharmaceuticals might be of interest in pulmonary or hepatic cancers, warranting further studies. The development of bivalent radiotracers to target several receptors concomitantly expressed could be of interest to improve targeting [237]. Similarly, improved detection and sensitivity could be achieved using bimodal agents [238]. Besides, the clinical success for radiolabeled somatostatin analogs both with diagnostic and therapeutic radionuclides paved the way for new promising peptide derivatives, such as bombesin, neurotensin, or CXCR4 ligands, and, in a similar way, PSMA ligands, for cancer theranostics [49,233,239,240].

Author Contributions

All authors contributed to the writing of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partly supported with a funding from Ligue Contre le Cancer (R.E.), a grant from Ligue 22 Contre le Cancer, and Labex IRON (grant no. ANR-11-LABX-0018).

Acknowledgments

The authors thank Sophie Laffont for providing the [68Ga]-DOTATOC and [177Lu]-DOTATATE pictures.

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. Patel, Y.C.; Greenwood, M.T.; Panetta, R.; Demchyshyn, L.; Niznik, H.; Srikant, C.B. The somatostatin receptor family. Life Sci. 1995, 57, 1249–1265. [Google Scholar] [CrossRef]
  3. 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] [Green Version]
  4. Patel, Y.C. Somatostatin and Its Receptor Family. Front. Neuroendocrinol. 1999, 20, 157–198. [Google Scholar] [CrossRef] [PubMed]
  5. Weckbecker, G.; Lewis, I.; Albert, R.; Schmid, H.A.; Hoyer, D.; Bruns, C. Opportunities in somatostatin research: Biological, chemical and therapeutic aspects. Nat. Rev. Drug Discov. 2003, 2, 999–1017. [Google Scholar] [CrossRef] [PubMed]
  6. Abdel-Rahman, O.; Lamarca, A.; Valle, J.W.; Hubner, R.A. Somatostatin receptor expression in hepatocellular carcinoma: Prognostic and therapeutic considerations. Endocr. Relat. Cancer 2014, 21, R485–R493. [Google Scholar] [CrossRef] [Green Version]
  7. Pyronnet, S.; Bousquet, C.; Najib, S.; Azar, R.; Laklai, H.; Susini, C. Antitumor effects of somatostatin. Mol. Cell. Endocrinol. 2008, 286, 230–237. [Google Scholar] [CrossRef] [Green Version]
  8. Barbieri, F.; Bajetto, A.; Pattarozzi, A.; Gatti, M.; Würth, R.; Thellung, S.; Corsaro, A.; Villa, V.; Nizzari, M.; Florio, T. Peptide receptor targeting in cancer: The somatostatin paradigm. Int. J. Pept. 2013, 2013, 926295. [Google Scholar] [CrossRef]
  9. 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]
  10. Reubi, J.C.; Schaer, J.C.; Laissue, J.A.; Waser, B. Somatostatin receptors and their subtypes in human tumors and in peritumoral vessels. Metabolism 1996, 45, 39–41. [Google Scholar] [CrossRef]
  11. 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] [PubMed]
  12. Guillermet-Guibert, J.; Lahlou, H.; Pyronnet, S.; Bousquet, C.; Susini, C. Somatostatin receptors as tools for diagnosis and therapy: Molecular aspects. Best Pract. Res. Clin. Gastroenterol. 2005, 19, 535551. [Google Scholar] [CrossRef]
  13. Hasskarl, J.; Kaufmann, M.; Schmid, H.A. Somatostatin receptors in non-neuroendocrine malignancies: The potential role of somatostatin analogs in solid tumors. Future Oncol. 2011, 7, 895–913. [Google Scholar] [CrossRef]
  14. 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] [Green Version]
  15. Hejna, M.; Schmidinger, M.; Raderer, M. The clinical role of somatostatin analogues as antineoplastic agents: Much ado about nothing? Ann. Oncol. 2002, 13, 653–668. [Google Scholar] [CrossRef]
  16. Keskin, O.; Yalcin, S. A review of the use of somatostatin analogs in oncology. OncoTargets Ther. 2013, 6, 471–483. [Google Scholar] [CrossRef] [Green Version]
  17. Reubi, J.C. Peptide receptors as molecular targets for cancer diagnosis and therapy. Endocr. Rev. 2003, 24, 389–427. [Google Scholar] [CrossRef] [Green Version]
  18. Zhao, B.; Zhao, H.; Zhao, N.; Zhu, X.G. Cholangiocarcinoma cells express somatostatin receptor subtype 2 and respond to octreotide treatment. J. Hepatobiliary Pancreat. Surg. 2002, 9, 497–502. [Google Scholar] [CrossRef]
  19. Bläker, M.; Schmitz, M.; Gocht, A.; Burghardt, S.; Schulz, M.; Bröring, D.C.; Pace, A.; Greten, H.; De Weerth, A. Differential expression of somatostatin receptor subtypes in hepatocellular carcinomas. J. Hepatol. 2004, 41, 112–118. [Google Scholar] [CrossRef]
  20. Reubi, J.C.; Waser, B. Concomitant expression of several peptide receptors in neuroendocrine tumours: Molecular basis for in vivo multireceptor tumour targeting. Eur. J. Nucl. Med. Mol. Imaging 2003, 30, 781–793. [Google Scholar] [CrossRef]
  21. Reubi, J.C.; Zimmermann, A.; Jonas, S.; Waser, B.; Neuhaus, P.; Läderach, U.; Wiedenmann, B. Regulatory peptide receptors in human hepatocellular carcinomas. Gut 1999, 45, 766–774. [Google Scholar] [CrossRef] [PubMed]
  22. 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] [PubMed] [Green Version]
  23. Nguyen-Khac, E.; Ollivier, I.; Aparicio, T.; Moullart, V.; Hugentobler, A.; Lebtahi, R.; Lobry, C.; Susini, C.; Duhamel, C.; Hommel, S.; et al. Somatostatin receptor scintigraphy screening in advanced hepatocarcinoma: A multi-center French study. Cancer Biol. Ther. 2009, 8, 2033–2039. [Google Scholar] [CrossRef] [PubMed]
  24. Verhoef, C.; van Dekken, H.; Hofland, L.J.; Zondervan, P.E.; de Wilt, J.H.; van Marion, R.; de Man, R.A.; IJzermans, J.N.; van Eijck, C.H. Somatostatin receptors in human hepatocellular carcinomas: Biological, patient and tumor characteristics. Dig. Surg. 2008, 25, 21–26. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. 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]
  26. Li, S.; Liu, Y.; Shen, Z. Characterization of somatostatin receptor 2 and 5 expression in operable hepatocellular carcinomas. Hepatogastroenterology 2012, 59, 2054–2058. [Google Scholar] [CrossRef]
  27. 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] [Green Version]
  28. Huang, C.Z.; Huang, A.M.; Liu, J.F.; Wang, B.; Lin, K.C.; Ye, Y.B. Somatostatin Octapeptide Inhibits Cell Invasion and Metastasis in Hepatocellular Carcinoma Through PEBP1. Cell Physiol. Biochem. 2018, 47, 2340–2349. [Google Scholar] [CrossRef]
  29. Hua, Y.P.; Huang, J.F.; Liang, L.J.; Li, S.Q.; Lai, J.M.; Liang, H.Z. The study of inhibition effect of octreotide on the growth of hepatocellular carcinoma xenografts in situ in nude mice. Chin. J. Surg. 2005, 43, 721–725. [Google Scholar]
  30. Jia, W.D.; Xu, G.L.; Wang, W.; Wang, Z.H.; Li, J.S.; Ma, J.L.; Ren, W.H.; Ge, Y.S.; Yu, J.H.; Liu, W.B. A somatostatin analogue, octreotide, inhibits the occurrence of second primary tumors and lung metastasis after resection of hepatocellular carcinoma in mice. Tohoku J. Exp. Med. 2009, 218, 155160. [Google Scholar] [CrossRef] [Green Version]
  31. 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] [Green Version]
  32. 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] [PubMed] [Green Version]
  33. 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]
  34. Becker, G.; Allgaier, H.P.; Olschewski, M.; Zähringer, A.; Blum, H.E.; HECTOR Study Group. Long-acting octreotide versus placebo for treatment of advanced HCC: A randomized controlled double-blind study. Hepatology 2007, 45, 9–15. [Google Scholar] [CrossRef]
  35. Barbare, J.C.; Bouché, 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 multicenter, randomised, double blind placebo-controlled study. Eur. J. Cancer 2009, 45, 1788–1797. [Google Scholar] [CrossRef]
  36. Samonakis, D.N.; Notas, G.; Christodoulakis, N.; Kouroumalis, E.A. Mechanisms of action and resistance of somatostatin analogues for the treatment of hepatocellular carcinoma: A message not well taken. Dig. Dis. Sci. 2008, 53, 2359–2365. [Google Scholar] [CrossRef]
  37. Kaemmerer, D.; Schindler, R.; Mußbach, F.; Dahmen, U.; Altendorf-Hofmann, A.; Dirsch, O.; Sänger, J.; Schulz, S.; Lupp, A. Somatostatin and CXCR4 chemokine receptor expression in hepatocellular and cholangiocellular carcinomas: Tumor capillaries as promising targets. BMC Cancer 2017, 17, 896. [Google Scholar] [CrossRef] [Green Version]
  38. Rinke, A.; Müller, H.H.; Schade-Brittinger, C.; Klose, K.J.; Barth, P.; Wied, M.; Mayer, C.; Aminossadati, B.; Pape, U.F.; Bläker, M.; et al. Placebo-controlled, double-blind, prospective, randomized study on the effect of octreotide LAR in the control of tumor growth in patients with metastatic neuroendocrine midgut tumors: A report from the PROMID study group. J. Clin. Oncol. 2009, 27, 4656–4663. [Google Scholar] [CrossRef]
  39. 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. Anti-tumour effects of lanreotide for pancreatic and intestinal neuroendocrine tumours: The CLARINET open-label extension study. Endocr. Relat. Cancer 2016, 23, 191–199. [Google Scholar] [CrossRef]
  40. Sheppard, M.; Shapiro, B.; Pimstone, B.; Kronheim, S.; Berelowitz, M.; Gregory, M. Metabolic clearance and plasma half-disappearance time of exogenous somatostatin in man. J. Clin. Endocrinol. Metab. 1979, 48, 50–53. [Google Scholar] [CrossRef]
  41. Lowell, A.; Freda, P.U. From somatostatin to octreotide LAR: Evolution of a somatostatin analogue. Curr. Med. Res. Opin. 2009, 25, 2989–2999. [Google Scholar] [CrossRef]
  42. Öberg, K.; Lamberts, S.W.J. Somatostatin analogues in acromegaly and gastroenteropancreatic neuroendocrine tumours: Past, present and future. Endocr. Relat. Cancer 2016, 23, R551–R566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Ryan, P.; McBride, A.; Ray, D.; Pulgar, S.; Ramirez, R.A.; Elquza, E.; Favaro, J.P.; Dranitsaris, G. Lanreotide vs. octreotide LAR for patients with advanced gastroenteropancreatic neuroendocrine tumors: An observational time and motion analysis. J. Oncol. Pharm. Pract. 2019, 25, 1425–1433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Feelders, R.A.; Yasothan, U.; Kirkpatrick, P. Pasireotide. Nat. Rev. Drug Discov. 2012, 11, 597–598. [Google Scholar] [CrossRef]
  45. Crider, A.M. Recent Advances in the Development of Nonpeptide Somatostatin Receptor Ligands. Mini Rev. Med. Chem. 2002, 2, 507–517. [Google Scholar] [CrossRef]
  46. Correia, J.D.; Paulo, A.; Raposinho, P.D.; Santos, I. Radiometallated peptides for molecular imaging and targeted therapy. Dalton Trans. 2011, 40, 6144–6167. [Google Scholar] [CrossRef]
  47. Jamous, M.; Haberkorn, U.; Mier, W. Synthesis of Peptide Radiopharmaceuticals for the Therapy and Diagnosis of Tumor Diseases. Molecules 2013, 18, 3379–3409. [Google Scholar] [CrossRef] [Green Version]
  48. Tornesello, A.L.; Buonaguro, L.; Tornesello, M.L.; Buonaguro, F.M. New Insights in the Design of Bioactive Peptides and Chelating Agents for Imaging and Therapy in Oncology. Molecules 2017, 22, 1282. [Google Scholar] [CrossRef] [Green Version]
  49. Tornesello, A.L.; Tornesello, M.L.; Buonaguro, F.M. An Overview of Bioactive Peptides for in vivo Imaging and Therapy in Human Diseases. Mini Rev. Med. Chem. 2017, 17, 758–770. [Google Scholar] [CrossRef]
  50. Cutler, C.S.; Hennkens, H.M.; Sisay, N.; Huclier-Markai, S.; Jurisson, S.S. Radiometals for Combined Imaging and Therapy. Chem. Rev. 2013, 113, 858–883. [Google Scholar] [CrossRef]
  51. Ramogida, C.F.; Orvig, C. Tumour Targeting with Radiometals for Diagnosis and Therapy. Chem. Commun. 2013, 49, 4720–4739. [Google Scholar] [CrossRef] [PubMed]
  52. Blower, P.J. A nuclear chocolate box: The periodic table of nuclear medicine. Dalton Trans. 2015, 44, 4819–4844. [Google Scholar] [CrossRef] [PubMed]
  53. Kostelnik, T.I.; Orvig, C. Radioactive Main Group and Rare Earth Metals for Imaging and Therapy. Chem. Rev. 2019, 119, 902–956. [Google Scholar] [CrossRef] [PubMed]
  54. Price, E.W.; Orvig, C. Matching Chelators to Radiometals for Radiopharmaceuticals. Chem. Soc. Rev. 2014, 43, 260–290. [Google Scholar] [CrossRef] [PubMed]
  55. Boros, E.; Packard, A.B. Radioactive Transition Metals for Imaging and Therapy. Chem. Rev. 2019, 119, 870–901. [Google Scholar] [CrossRef]
  56. Hancock, R.D.; Martell, A.E. Ligand design for selective complexation of metal ions in aqueous solution. Chem. Rev. 1989, 89, 1875–1914. [Google Scholar] [CrossRef]
  57. Stasiuk, G.J.; Long, N.J. The ubiquitous DOTA and its derivatives: The impact of 1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetic acid on biomedical imaging. Chem. Commun. 2013, 49, 2732–2746. [Google Scholar] [CrossRef]
  58. Baranyai, Z.; Tircsó, G.; Rösch, F. The Use of the Macrocyclic Chelator DOTA in Radiochemical Separations. Eur. J. Inorg. Chem. 2020, 36–56. [Google Scholar] [CrossRef] [Green Version]
  59. Sun, X.; Wuest, M.; Weisman, G.R.; Wong, E.H.; Reed, D.P.; Boswell, C.A.; Motekaitis, R.; Martell, A.E.; Welch, M.J.; Anderson, C.J. Radiolabeling and in vivo behavior of copper-64-labeled cross-bridged cyclam ligands. J. Med. Chem. 2002, 45, 469–477. [Google Scholar] [CrossRef]
  60. Bhattacharyya, S.; Dixit, M. Metallic radionuclides in the development of diagnostic and therapeutic radiopharmaceuticals. Dalton Trans. 2011, 40, 6112–6128. [Google Scholar] [CrossRef]
  61. Mushtaq, S.; Yun, S.J.; Jeon, J. Recent Advances in Bioorthogonal Click Chemistry for Efficient Synthesis of Radiotracers and Radiopharmaceuticals. Molecules 2019, 24, 3567. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Reubi, J.C.; Maecke, H.R. Peptide-based probes for cancer imaging. J. Nucl. Med. 2008, 49, 1735–1738. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Maecke, H.R.; Reubi, J.C. Somatostatin receptors as targets for nuclear medicine imaging and radionuclide treatment. J. Nucl. Med. 2011, 52, 841–844. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Reubi, J.C.; Schär, J.C.; Waser, B.; Wenger, S.; Heppeler, A.; Schmitt, J.S.; Mäcke, H.R. Affinity profiles for human somatostatin receptor subtypes SST1-SST5 of somatostatin radiotracers selected for scintigraphic and radiotherapeutic use. Eur. J. Nucl. Med. 2000, 27, 273–282. [Google Scholar] [CrossRef]
  65. Fani, M.; Maecke, H.R. Radiopharmaceutical development of radiolabeled peptides. Eur. J. Nucl. Med. Mol. Imaging 2012, 39, S11–S30. [Google Scholar] [CrossRef]
  66. Lamberts, S.W.; Bakker, W.H.; Reubi, J.C.; Krenning, E.P. Somatostatin-receptor imaging in the localization of endocrine tumors. N. Engl. J. Med. 1990, 323, 1246–1249. [Google Scholar] [CrossRef]
  67. Bakker, W.H.; Krenning, E.P.; Breeman, W.A.; Kooij, P.P.; Reubi, J.C.; Koper, J.W.; de Jong, M.; Laméris, J.S.; Visser, T.J.; Lamberts, S.W. In vivo use of a radioiodinated somatostatin analogue: Dynamics, metabolism, and binding to somatostatin receptor-positive tumors in man. J. Nucl. Med. 1991, 32, 1184–1189. [Google Scholar]
  68. Bakker, W.H.; Krenning, E.P.; Breeman, W.A.; Koper, J.W.; Kooij, P.P.; Reubi, J.C.; Klijn, J.G.; Visser, T.J.; Docter, R.; Lamberts, S.W. Receptor scintigraphy with a radioiodinated somatostatin analogue: Radiolabeling, purification, biologic activity, and in vivo application in animals. J. Nucl. Med. 1990, 31, 1501–1509. [Google Scholar]
  69. Bakker, W.H.; Albert, R.; Bruns, C.; Breeman, W.A.; Hofland, L.J.; Marbach, P.; Pless, J.; Pralet, D.; Stolz, B.; Koper, J.W.; et al. [111In-DTPA-D-Phe1]-octreotide, a potential radiopharmaceutical for imaging of somatostatin receptor-positive tumors: Synthesis, radiolabeling and in vitro validation. Life Sci. 1991, 49, 1583–1591. [Google Scholar] [CrossRef]
  70. Bakker, W.H.; Krenning, E.P.; Reubi, J.C.; Breeman, W.A.; Setyono-Han, B.; de Jong, M.; Kooij, P.P.; Bruns, C.; van Hagen, P.M.; Marbach, P.; et al. In vivo application of [111In-DTPA-D-Phe1]-octreotide for detection of somatostatin receptor-positive tumors in rats. Life Sci. 1991, 49, 1593–1601. [Google Scholar] [CrossRef]
  71. Krenning, E.P.; Bakker, W.H.; Kooij, P.P.; Breeman, W.A.; Oei, H.Y.; de Jong, M.; Reubi, J.C.; Visser, T.J.; Bruns, C.; Kwekkeboom, D.J.; et al. Somatostatin receptor scintigraphy with indium-111-DTPA-D-Phe-1-octreotide in man: Metabolism, dosimetry and comparison with iodine-123-Tyr-3-octreotide. J. Nucl. Med. 1992, 33, 652–658. [Google Scholar] [PubMed]
  72. Otte, A.; Jermann, E.; Behe, M.; Goetze, M.; Bucher, H.C.; Roser, H.W.; Heppeler, A.; Mueller-Brand, J.; Maecke, H.R. DOTATOC: A powerful new tool for receptor-mediated radionuclide therapy. Eur. J. Nucl. Med. 1997, 24, 792–795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Forrer, F.; Uusijärvi, H.; Waldherr, C.; Cremonesi, M.; Bernhardt, P.; Mueller-Brand, J.; Maecke, H.R. A comparison of 111In-DOTATOC and 111In-DOTATATE: Biodistribution and dosimetry in the same patients with metastatic neuroendocrine tumours. Eur. J. Nucl. Med. Mol. Imaging 2004, 31, 1257–1262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Wild, D.; Schmitt, J.S.; Ginj, M.; Mäcke, H.R.; Bernard, B.F.; Krenning, E.; De Jong, M.; Wenger, S.; Reubi, J.C. DOTA-NOC, a high-affinity ligand of somatostatin receptor subtypes 2, 3 and 5 for labeling with various radiometals. Eur. J. Nucl. Med. Mol. Imaging 2003, 30, 1338–1347. [Google Scholar] [CrossRef]
  75. Virgolini, I.; Ambrosini, V.; Bomanji, J.B.; Baum, R.P.; Fanti, S.; Gabriel, M.; Papathanasiou, N.D.; Pepe, G.; Oyen, W.; De Cristoforo, C.; et al. Procedure guidelines for PET/CT tumour imaging with 68Ga-DOTA-conjugated peptides: 68Ga-DOTA-TOC, 68Ga-DOTA-NOC, 68Ga-DOTA-TATE. Eur. J. Nucl. Med. Mol. Imaging 2010, 37, 2004–2010. [Google Scholar] [CrossRef]
  76. Hofmann, M.; Maecke, H.; Börner, R.; Weckesser, E.; Schöffski, P.; Oei, L.; Schumacher, J.; Henze, M.; Heppeler, A.; Meyer, J.; et al. Biokinetics and imaging with the somatostatin receptor PET radioligand 68Ga-DOTATOC: Preliminary data. Eur. J. Nucl. Med. 2001, 28, 1751–1757. [Google Scholar] [CrossRef]
  77. Hofman, M.S.; Lau, W.F.; Hicks, R.J. Somatostatin receptor imaging with 68Ga DOTATATE PET/CT: Clinical utility, normal patterns, pearls, and pitfalls in interpretation. Radiographics 2015, 35, 500–516. [Google Scholar] [CrossRef] [Green Version]
  78. Wild, D.; Mäcke, H.R.; Waser, B.; Reubi, J.C.; Ginj, M.; Rasch, H.; Müller-Brand, J.; Hofmann, M. 68Ga-DOTANOC: A first compound for PET imaging with high affinity for somatostatin receptor subtypes 2 and 5. Eur. J. Nucl. Med. Mol. Imaging 2005, 32, 724. [Google Scholar] [CrossRef]
  79. Antunes, P.; Ginj, M.; Walter, M.A.; Chen, J.; Reubi, J.C.; Maecke, H.R. Influence of different spacers on the biological profile of a DOTA-Somatostatin analogue. Bioconjugate Chem. 2007, 18, 84–92. [Google Scholar] [CrossRef]
  80. Eisenwiener, K.P.; Prata, M.I.; Buschmann, I.; Zhang, H.W.; Santos, A.C.; Wenger, S.; Reubi, J.C.; Mäcke, H.R. NODAGATOC, a new chelator-coupled somatostatin analogue labeled with [67/68Ga] and [111In] for SPECT, PET, and targeted therapeutic applications of somatostatin receptor (hsst2) expressing tumors. Bioconjugate Chem. 2002, 13, 530–541. [Google Scholar] [CrossRef]
  81. Laznickova, A.; Laznicek, M.; Trejtnar, F.; Maecke, H.R.; Eisenwiener, K.P.; Reubi, J.C. Biodistribution of two octreotate analogs radiolabeled with indium and yttrium in rats. Anticancer Res. 2010, 30, 2177–2184. [Google Scholar] [PubMed]
  82. Ginj, M.; Chen, J.; Walter, M.A.; Eltschinger, V.; Reubi, J.C.; Maecke, H.R. Preclinical evaluation of new and highly potent analogues of octreotide for predictive imaging and targeted radiotherapy. Clin. Cancer Res. 2005, 11, 1136–1145. [Google Scholar] [PubMed]
  83. Boubaker, A.; Prior, J.O.; Willi, J.P.; Champendal, M.; Kosinski, M.; Bischof-Delaloye, A.; Maecke, H.R.; Ginj, M.; Baechler, S.; Buchegger, F. Biokinetics and dosimetry of 111In-DOTA-NOC-ATE compared with 111In-DTPA-octreotide. Eur. J. Nucl. Med. Mol. Imaging 2012, 39, 1868–1875. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Ma, M.T.; Cullinane, C.; Waldeck, K.; Roselt, P.; Hicks, R.J.; Blower, P.J. Rapid kit-based 68Ga-labeling and PET imaging with THP-Tyr3-octreotate: A preliminary comparison with DOTA-Tyr3octreotate. EJNMMI Res. 2015, 5, 52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Fani, M.; Mueller, A.; Tamma, M.L.; Nicolas, G.; Rink, H.R.; Cescato, R.; Reubi, J.C.; Maecke, H.R. Radiolabeled bicyclic somatostatin-based analogs: A novel class of potential radiotracers for SPECT/PET of neuroendocrine tumors. J. Nucl. Med. 2010, 51, 1771–1779. [Google Scholar] [CrossRef] [Green Version]
  86. Ginj, M.; Zhang, H.; Eisenwiener, K.P.; Wild, D.; Schulz, S.; Rink, H.; Cescato, R.; Reubi, J.C.; Maecke, H.R. New pansomatostatin ligands and their chelated versions: Affinity profile, agonist activity, internalization, and tumor targeting. Clin. Cancer Res. 2008, 14, 2019–2027. [Google Scholar] [CrossRef] [Green Version]
  87. Liu, F.; Liu, T.; Xu, X.; Guo, X.; Li, N.; Xiong, C.; Li, C.; Zhu, H.; Yang, Z. Design, Synthesis, and Biological Evaluation of 68Ga-DOTA-PA1 for Lung Cancer: A Novel PET Tracer for Multiple Somatostatin Receptor Imaging. Mol. Pharm. 2018, 15, 619–628. [Google Scholar] [CrossRef]
  88. Tatsi, A.; Maina, T.; Cescato, R.; Waser, B.; Krenning, E.P.; de Jong, M.; Cordopatis, P.; Reubi, J.C.; Nock, B.A. [111In-DOTA]Somatostatin-14 analogs as potential pansomatostatin-like radiotracers—First results of a preclinical study. EJNMMI Res. 2012, 2, 25. [Google Scholar] [CrossRef] [Green Version]
  89. Maina, T.; Cescato, R.; Waser, B.; Tatsi, A.; Kaloudi, A.; Krenning, E.P.; de Jong, M.; Nock, B.A.; Reubi, J.C. [111In-DOTA]LTT-SS28, a first pansomatostatin radioligand for in vivo targeting of somatostatin receptor-positive tumors. J. Med. Chem. 2014, 57, 6564–6571. [Google Scholar] [CrossRef]
  90. Pearson, D.A.; Lister-James, J.; McBride, W.J.; Wilson, D.M.; Martel, L.J.; Civitello, E.R.; Taylor, J.E.; Moyer, B.R.; Dean, R.T. Somatostatin receptor-binding peptides labeled with technetium-99m: Chemistry and initial biological studies. J. Med. Chem. 1996, 39, 1361–1371. [Google Scholar] [CrossRef]
  91. Decristoforo, C.; Mather, S.J. Preparation, 99mTc-labeling and in vitro characterization of HYNIC and N3S modified RC-160 and [Tyr3]octreotide. Bioconjugate Chem. 1999, 10, 431–438. [Google Scholar] [CrossRef] [PubMed]
  92. Maina, T.; Stolz, B.; Albert, R.; Bruns, C.; Koch, P.; Mäcke, H. Synthesis, radiochemistry and biological evaluation of a new somatostatin analogue (SDZ 219-387) labeled with technetium-99m. Eur. J. Nucl. Med. 1994, 21, 437–444. [Google Scholar] [CrossRef] [PubMed]
  93. Thakur, M.L.; Kolan, H.; Li, J.; Wiaderkiewicz, R.; Pallela, V.R.; Duggaraju, R.; Schally, A.V. Radiolabeled somatostatin analogues in prostate cancer. Nucl. Med. Biol. 1997, 24, 105–113. [Google Scholar] [CrossRef]
  94. Abiraj, K.; Ursillo, S.; Tamma, M.L.; Rylova, S.N.; Waser, B.; Constable, E.C.; Fani, M.; Nicolas, G.P.; Reubi, J.C.; Maecke, H.R. The tetraamine chelator outperforms HYNIC in a new technetium-99m-labeled somatostatin receptor 2 antagonist. EJNMMI Res. 2018, 8, 75. [Google Scholar] [CrossRef]
  95. Spradau, T.W.; Edwards, W.B.; Anderson, C.J.; Welch, M.J.; Katzenellenbogen, J.A. Synthesis and biological evaluation of Tc-99m-cyclopentadienyltricarbonyltechnetium-labeled octreotide. Nucl. Med. Biol. 1999, 26, 1–7. [Google Scholar] [CrossRef]
  96. Makris, G.; Kuchuk, M.; Gallazzi, F.; Jurisson, S.S.; Smith, C.J.; Hennkens, H.M. Somatostatin receptor targeting with hydrophilic [99mTc/186Re]Tc/Re-tricarbonyl NODAGA and NOTA complexes. Nucl. Med. Biol. 2019, 71, 39–46. [Google Scholar] [CrossRef]
  97. Abrams, M.J.; Juweid, M.; tenKate, C.I.; Schwartz, D.A.; Hauser, M.M.; Gaul, F.E.; Fuccello, A.J.; Rubin, R.H.; Strauss, H.W.; Fischman, A.J. Technetium-99m-human polyclonal IgG radiolabeled via the hydrazino nicotinamide derivative for imaging focal sites of infection in rats. J. Nucl. Med. 1990, 31, 2022–2028. [Google Scholar]
  98. Mikolajczak, R.; Maecke, H.R. Radiopharmaceuticals for somatostatin receptor imaging. Nucl. Med. Rev. Cent. East Eur. 2016, 19, 126–132. [Google Scholar] [CrossRef] [Green Version]
  99. Decristoforo, C.; Melendez-Alafort, L.; Sosabowski, J.K.; Mather, S.J. 99mTc-HYNIC-[Tyr3]octreotide for imaging somatostatin receptor positive tumors: Preclinical evaluation and comparison with 111In-octreotide. J. Nucl. Med. 2000, 41, 1114–1119. [Google Scholar]
  100. Gabriel, M.; Decristoforo, C.; Donnemiller, E.; Ulmer, H.; Watfah Rychlinski, C.; Mather, S.J.; Moncayo, R. An intrapatient comparison of 99mTc-EDDA/HYNIC-TOC with 111In-DTPA-Octreotide for diagnosis of somatostatin receptor expressing tumors. J. Nucl. Med. 2003, 44, 708–716. [Google Scholar]
  101. Cwikla, J.B.; Mikolajczak, R.; Pawlak, D.; Buscombe, J.R.; Nasierowska-Guttmejer, A.; Bator, A.; Maecke, H.R.; Walecki, J. Initial direct comparison of 99mTc-TOC and 99mTc-TATE in identifying sites of disease in patients with proven GEP NETs. J. Nucl. Med. 2008, 49, 1060–1065. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Decristoforo, C.; Maina, T.; Nock, B.; Gabriel, M.; Cordopatis, P.; Moncayo, R. 99mTc-Demotate 1: First data in tumour patients-results of a pilot/phase I study. Eur. J. Nucl. Med. Mol. Imaging 2003, 30, 1211–1219. [Google Scholar] [CrossRef] [PubMed]
  103. Maina, T.; Nock, B.A.; Cordopatis, P.; Bernard, B.F.; Breeman, W.A.; van Gameren, A.; van den Berg, R.; Reubi, J.C.; Krenning, E.P.; de Jong, M. [99mTc]Demotate 2 in the detection of sst2-positive tumours: A preclinical comparison with [111In]DOTA-tate. Eur. J. Nucl. Med. Mol. Imaging 2006, 33, 831–840. [Google Scholar] [CrossRef] [PubMed]
  104. Virgolini, I.; Leimer, M.; Handmaker, H.; Lastoria, S.; Bischof, C.; Muto, P.; Pangerl, T.; Gludovacz, D.; Peck-Radosavljevic, M.; Lister-James, J.; et al. Somatostatin receptor subtype specificity and in vivo binding of a novel tumor tracer, 99mTc-P829. Cancer Res. 1998, 58, 1850–1859. [Google Scholar]
  105. Lebtahi, R.; Le Cloirec, J.; Houzard, C.; Daou, D.; Sobhani, I.; Sassolas, G.; Mignon, M.; Bourguet, P.; Le Guludec, D. Detection of neuroendocrine tumors: 99mTc-P829 scintigraphy compared with 111In-pentetreotide scintigraphy. J. Nucl. Med. 2002, 43, 889–895. [Google Scholar]
  106. Blum, J.E.; Handmaker, H.; Rinne, N.A. The utility of a somatostatin-type receptor binding peptide radiopharmaceutical (P829) in the evaluation of solitary pulmonary nodules. Chest 1999, 115, 224–232. [Google Scholar] [CrossRef] [Green Version]
  107. Bååth, M.; Kolbeck, K.G.; Danielsson, R. Somatostatin receptor scintigraphy with 99mTc-Depreotide (NeoSpect) in discriminating between malignant and benign lesions in the diagnosis of lung cancer: A pilot study. Acta Radiol. 2004, 45, 833–839. [Google Scholar] [CrossRef]
  108. Axelsson, R.; Herlin, G.; Bååth, M.; Aspelin, P.; Kolbeck, K.G. Role of scintigraphy with technetium-99m depreotide in the diagnosis and management of patients with suspected lung cancer. Acta Radiol. 2008, 49, 295–302. [Google Scholar] [CrossRef]
  109. Menda, Y.; Kahn, D. Somatostatin receptor imaging of non-small cell lung cancer with 99mTc depreotide. Semin. Nucl. Med. 2002, 32, 92–96. [Google Scholar] [CrossRef]
  110. Van Den Bossche, B.; D’haeninck, E.; Bacher, K.; Thierens, H.; Van Belle, S.; Dierckx, R.A.; Van de Wiele, C. Biodistribution and dosimetry of 99mTc-depreotide (P829) in patients suffering from breast carcinoma. Cancer Biother. Radiopharm. 2004, 19, 776–783. [Google Scholar] [CrossRef]
  111. Briganti, V.; Cuccurullo, V.; Di Stasio, G.D.; Mansi, L. Gamma Emitters in Pancreatic Endocrine Tumors Imaging in the PET Era: Is there a Clinical Space for 99mTc-peptides? Curr. Radiopharm. 2019, 12, 156–170. [Google Scholar] [CrossRef] [PubMed]
  112. Boschi, A.; Uccelli, L.; Martini, P. A Picture of Modern Tc-99m Radiopharmaceuticals: Production, Chemistry, and Applications in Molecular Imaging. Appl. Sci. 2019, 9, 2526. [Google Scholar] [CrossRef] [Green Version]
  113. Boschi, A.; Martini, P.; Janevik-Ivanovska, E.; Duatti, A. The emerging role of copper-64 radiopharmaceuticals as cancer theranostics. Drug Discov. Today 2018, 23, 1489–1501. [Google Scholar] [CrossRef] [PubMed]
  114. Anderson, C.J.; Dehdashti, F.; Cutler, P.D.; Schwarz, S.W.; Laforest, R.; Bass, L.A.; Lewis, J.S.; McCarthy, D.W. 64Cu-TETA-octreotide as a PET imaging agent for patients with neuroendocrine tumors. J. Nucl. Med. 2001, 42, 213–221. [Google Scholar]
  115. Sprague, J.E.; Peng, Y.; Sun, X.; Weisman, G.R.; Wong, E.H.; Achilefu, S.; Anderson, C.J. Preparation and biological evaluation of copper-64-labeled Tyr3-octreotate using a cross-bridged macrocyclic chelator. Clin. Cancer Res. 2004, 10, 8674–8682. [Google Scholar] [CrossRef] [Green Version]
  116. Edwards, W.B.; Fields, C.G.; Anderson, C.J.; Pajeau, T.S.; Welch, M.J.; Fields, G.B. Generally Applicable, Convenient Solid-Phase Synthesis and Receptor Affinities of Octreotide Analogs. J. Med. Chem. 1994, 37, 3749–3757. [Google Scholar] [CrossRef]
  117. Paterson, B.M.; Roselt, P.; Denoyer, D.; Cullinane, C.; Binns, D.; Noonan, W.; Jeffery, C.M.; Price, R.I.; White, J.M.; Hicks, R.J.; et al. PET imaging of tumours with a 64Cu labeled macrobicyclic cage amine ligand tethered to Tyr3-octreotate. Dalton Trans. 2014, 43, 1386–1396. [Google Scholar] [CrossRef]
  118. Marciniak, A.; Brasuń, J. Somatostatin analogues labeled with copper radioisotopes: Currrent status. J. Radioanal. Nucl. Chem. 2017, 313, 279–289. [Google Scholar] [CrossRef]
  119. Pfeifer, A.; Knigge, U.; Binderup, T.; Mortensen, J.; Oturai, P.; Loft, A.; Berthelsen, A.K.; Langer, S.W.; Rasmussen, P.; Elema, D.; et al. 64Cu-DOTATATE PET for Neuroendocrine Tumors: A Prospective Head-to-Head Comparison with 111In-DTPA-Octreotide in 112 Patients. J. Nucl. Med. 2015, 56, 847–854. [Google Scholar] [CrossRef] [Green Version]
  120. Johnbeck, C.B.; Knigge, U.; Loft, A.; Berthelsen, A.K.; Mortensen, J.; Oturai, P.; Langer, S.W.; Elema, D.R.; Kjaer, A. Head-to-head comparison of 64Cu-DOTATATE and 68Ga-DOTATOC PET/CT: A prospective study of 59 patients with neuroendocrine tumors. J. Nucl. Med. 2017, 58, 451–458. [Google Scholar] [CrossRef] [Green Version]
  121. Carlsen, E.A.; Johnbeck, C.B.; Binderup, T.; Loft, M.; Pfeifer, A.; Mortensen, J.; Oturai, P.; Loft, A.; Berthelsen, A.K.; Langer, S.W.; et al. 64Cu-DOTATATE PET/CT and prediction of overall and progression-free survival in patients with neuroendocrine neoplasms. J. Nucl. Med. 2020. [Google Scholar] [CrossRef]
  122. Andersen, T.L.; Baun, C.; Olsen, B.B.; Dam, J.H.; Thisgaard, H. Improving Contrast and Detectability: Imaging with [55Co]Co-DOTATATE in Comparison with [64Cu]Cu-DOTATATE and [68Ga]Ga-DOTATATE. J. Nucl. Med. 2020, 61, 228–233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Mirzaei, S.; Revheim, M.; Raynor, W.; Zehetner, W.; Knoll, P.; Zandieh, S.; Alavi, A. 64Cu-DOTATOC PET-CT in Patients with Neuroendocrine Tumors. Oncol. Ther. 2019, 8, 125–131. [Google Scholar] [CrossRef] [Green Version]
  124. Hicks, R.J.; Jackson, P.; Kong, G.; Ware, R.E.; Hofman, M.S.; Pattison, D.A.; Akhurst, T.A.; Drummond, E.; Roselt, P.; Callahan, J.; et al. 64Cu-SARTATE PET Imaging of Patients with Neuroendocrine Tumors Demonstrates High Tumor Uptake and Retention, Potentially Allowing Prospective Dosimetry for Peptide Receptor Radionuclide Therapy. J. Nucl. Med. 2019, 60, 777–785. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Delpassand, E.S.; Ranganathan, D.; Wagh, N.; Shafie, A.; Gaber, A.; Abbasi, A.; Kjaer, A.; Tworowska, I.; Núñez, R. 64Cu-DOTATATE PET/CT for Imaging Patients with Known or Suspected Somatostatin Receptor-Positive Neuroendocrine Tumors: Results of the First US Prospective, Reader-Blinded Clinical Trial. J. Nucl. Med. 2020, 61, 890–896. [Google Scholar] [CrossRef] [PubMed]
  126. Heppeler, A.; André, J.P.; Buschmann, I.; Wang, X.; Reubi, J.C.; Hennig, M.; Kaden, T.A.; Maecke, H.R. Metal-ion-dependent biological properties of a chelator-derived somatostatin analogue for tumour targeting. Chem. Eur. J. 2008, 14, 3026–3034. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Thisgaard, H.; Olsen, B.B.; Dam, J.H.; Bollen, P.; Mollenhauer, J.; Høilund-Carlsen, P.F. Evaluation of cobalt-labeled octreotide analogs for molecular imaging and Auger electron-based radionuclide therapy. J. Nucl. Med. 2014, 55, 1311–1316. [Google Scholar] [CrossRef] [Green Version]
  128. Thisgaard, H.; Olesen, M.; Dam, J.H. Radiosynthesis of Co-55- and Co-58m-labeled DOTATOC for positron emission tomography imaging and targeted radionuclide therapy. J. Label. Compd. Radiopharm. 2011, 54, 758–762. [Google Scholar] [CrossRef]
  129. Müller, C.; Domnanich, K.A.; Umbricht, C.A.; van der Meulen, N.P. Scandium and terbium radionuclides for radiotheranostics: Current state of development towards clinical application. Br. J. Radiol. 2018, 91, 20180074. [Google Scholar] [CrossRef]
  130. Pruszyński, M.; Majkowska-Pilip, A.; Loktionova, N.S.; Eppard, E.; Roesch, F. Radiolabeling of DOTATOC with the long-lived positron emitter 44Sc. Appl. Radiat. Isot. 2012, 70, 974–979. [Google Scholar] [CrossRef]
  131. Müller, C.; Vermeulen, C.; Johnston, K.; Köster, U.; Schmid, R.; Türler, A.; van der Meulen, N.P. Preclinical in vivo application of 152Tb-DOTANOC: A radiolanthanide for PET imaging. EJNMMI Res. 2016, 6, 35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Singh, A.; van der Meulen, N.P.; Muller, C.; Klette, I.; Kulkarni, H.R.; Türler, A.; Schibli, R.; Baum, R.P. First-in-human PET/CT imaging of metastatic neuroendocrine neoplasms with cyclotron-produced 44Sc-DOTATOC: A proof-of-concept study. Cancer Biother. Radiopharm. 2017, 32, 124–132. [Google Scholar] [CrossRef] [PubMed]
  133. Baum, R.P.; Singh, A.; Benešová, M.; Vermeulen, C.; Gnesin, S.; Köster, U.; Johnston, K.; Müller, D.; Senftleben, S.; Kulkarni, H.R.; et al. Clinical evaluation of the radiolanthanide terbium-152: First-in-human PET/CT with 152Tb-DOTATOC. Dalton Trans. 2017, 46, 14638–14646. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Koumarianou, E.; Pawlak, D.; Korsak, A.; Mikolajczak, R. Comparison of receptor affinity of natSc-DOTA-TATE versus natGa-DOTA-TATE. Nucl. Med. Rev. Cent. East Eur. 2011, 14, 85–89. [Google Scholar] [CrossRef]
  135. Domnanich, K.A.; Müller, C.; Farkas, R.; Schmid, R.M.; Ponsard, B.; Schibli, R.; Türler, A.; van der Meulen, N.P. 44Sc for labeling of DOTA- and NODAGA-functionalized peptides: Preclinical in vitro and in vivo investigations. EJNMMI Radiopharm. Chem. 2017, 1, 8. [Google Scholar] [CrossRef] [Green Version]
  136. Sinnes, J.P.; Nagel, J.; Rösch, F. AAZTA5/AAZTA5-TOC: Synthesis and radiochemical evaluation with 68Ga, 44Sc and 177Lu. EJNMMI Radiopharm. Chem. 2019, 4, 18. [Google Scholar] [CrossRef] [Green Version]
  137. Müller, C.; Fischer, E.; Behe, M.; Köster, U.; Dorrer, H.; Reber, J.; Haller, S.; Cohrs, S.; Blanc, A.; Grünberg, J.; et al. Future prospects for SPECT imaging using the radiolanthanide terbium-155—Production and preclinical evaluation in tumor-bearing mice. Nucl. Med. Biol. 2014, 41, e58–e65. [Google Scholar] [CrossRef]
  138. Walrand, S.; Jamar, F.; Mathieu, I.; De Camps, J.; Lonneux, M.; Sibomana, M.; Labar, D.; Michel, C.; Pauwels, S. Quantitation in PET using isotopes emitting prompt single gammas: Application to yttrium-86. Eur. J. Nucl. Med. Mol. Imaging 2003, 30, 354–361. [Google Scholar] [CrossRef]
  139. Helisch, A.; Förster, G.J.; Reber, H.; Buchholz, H.G.; Arnold, R.; Göke, B.; Weber, M.M.; Wiedenmann, B.; Pauwels, S.; Haus, U.; et al. Pre-therapeutic dosimetry and biodistribution of 86Y-DOTA-Phe1-Tyr3-octreotide versus 111In-pentetreotide in patients with advanced neuroendocrine tumours. Eur. J. Nucl. Med. Mol. Imaging 2004, 31, 1386–1392. [Google Scholar] [CrossRef]
  140. Clifford, T.; Boswell, C.A.; Biddlecombe, G.B.; Lewis, J.S.; Brechbiel, M.W. Validation of a novel CHX-A “derivative suitable for peptide conjugation: Small animal PET/CT imaging using yttrium-86-CHX-A”-octreotide. J. Med. Chem. 2006, 49, 4297–4304. [Google Scholar] [CrossRef]
  141. Jamar, F.; Barone, R.; Mathieu, I.; Walrand, S.; Labar, D.; Carlier, P.; de Camps, J.; Schran, H.; Chen, T.; Smith, M.C.; et al. 86Y-DOTA0-D-Phe1-Tyr3-octreotide (SMT487)—A phase 1 clinical study: Pharmacokinetics, biodistribution and renal protective effect of different regimens of amino acid co-infusion. Eur. J. Nucl. Med. Mol. Imaging 2003, 30, 510–518. [Google Scholar] [CrossRef] [PubMed]
  142. Mikolajczak, R.; van der Meulen, N.P.; Lapi, S.E. Radiometals for imaging and theranostics, current production, and future perspectives. J. Label. Compd. Radiopharm. 2019, 62, 615–634. [Google Scholar] [CrossRef] [PubMed]
  143. do Carmo, S.J.C.; Scott, P.J.H.; Alves, F. Production of radiometals in liquid targets. EJNMMI Radiopharm. Chem. 2020, 5, 2. [Google Scholar] [CrossRef] [PubMed]
  144. Talip, Z.; Favaretto, C.; Geistlich, S.; Meulen, N.P.V. A Step-by-Step Guide for the Novel Radiometal Production for Medical Applications: Case Studies with 68Ga, 44Sc, 177Lu and 161Tb. Molecules 2020, 25, 966. [Google Scholar] [CrossRef] [Green Version]
  145. Waldmann, C.M.; Stuparu, A.D.; van Dam, R.M.; Slavik, R. The search for an alternative to [68Ga]Ga-DOTA-TATE in neuroendocrine tumor theranostics: Current state of 18F-labeled somatostatin analog development. Theranostics 2019, 9, 1336–1347. [Google Scholar] [CrossRef]
  146. Guhlke, S.; Wester, H.J.; Bruns, C.; Stöcklin, G. (2-[18F]Fluoropropionyl-(D)phe1)-octreotide, a potential radiopharmaceutical for quantitative somatostatin receptor imaging with PET: Synthesis, Radiolabeling, in vitro validation and biodistribution in mice. Nucl. Med. Biol. 1994, 21, 819–825. [Google Scholar] [CrossRef]
  147. Hostetler, E.D.; Edwards, W.B.; Anderson, C.J.; Welch, M.J. Synthesis of 4-[18F]fluorobenzoyl octreotide and biodistribution in tumour-bearing Lewis rats. J. Label. Compd Radiopharm. 1999, 42, S720–S722. [Google Scholar]
  148. Wester, H.J.; Schottelius, M.; Poethko, T.; Bruus-Jensen, K.; Schwaiger, M. Radiolabeled Carbohydrated Somatostatin Analogs: A Review of the Current Status. Cancer Biother. Radiopharm. 2004, 19, 231–244. [Google Scholar] [CrossRef]
  149. Maschauer, S.; Heilmann, M.; Wängler, C.; Schirrmacher, R.; Prante, O. Radiosynthesis and preclinical evaluation of 18F-fluoroglycosylated octreotate for somatostatin receptor imaging. Bioconjugate Chem. 2016, 27, 2707–2714. [Google Scholar] [CrossRef]
  150. Liu, Z.; Pourghiasian, M.; Bénard, F.; Pan, J.; Lin, K.S.; Perrin, D.M. Preclinical evaluation of a high-affinity 18F-trifluoroborate octreotate derivative for somatostatin receptor imaging. J. Nucl. Med. 2014, 55, 1499–1505. [Google Scholar] [CrossRef] [Green Version]
  151. Niedermoser, S.; Chin, J.; Wängler, C.; Kostikov, A.; Bernard-Gauthier, V.; Vogler, N.; Soucy, J.P.; McEwan, A.J.; Schirrmacher, R.; Wängler, B. In vivo evaluation of 18F-SiFAlin-Modified TATE: A potential challenge for 68Ga-DOTATATE, the clinical gold standard for somatostatin receptor imaging with PET. J. Nucl. Med. 2015, 56, 1100–1105. [Google Scholar] [CrossRef] [Green Version]
  152. Allott, L.; Dubash, S.; Aboagye, E.O. [18F]FET-βAG-TOCA: The Design, Evaluation and Clinical Translation of a Fluorinated Octreotide. Cancers 2020, 12, 865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Laverman, P.; McBride, W.J.; Sharkey, R.M.; Eek, A.; Joosten, L.; Oyen, W.J.G.; Goldenberg, D.M.; Boerman, O.C. A novel facile method of labeling octreotide with 18F-fluorine. J. Nucl. Med. 2010, 51, 454–461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Meisetschlaeger, G.; Poethko, T.; Stahl, A.; Wolf, I.; Scheidhauer, K.; Schottelius, M.; Herz, M.; Wester, H.J.; Schwaiger, M. Gluc-Lys([18F]FP)-TOCA PET in patients with SSTR-positive tumors: Biodistribution and diagnostic evaluation compared with [111In]DTPA-octreotide. J. Nucl. Med. 2006, 47, 566–573. [Google Scholar]
  155. Ilhan, H.; Lindner, S.; Todica, A.; Cyran, C.C.; Tiling, R.; Auernhammer, C.J.; Spitzweg, C.; Boeck, S.; Unterrainer, M.; Gildehaus, F.J.; et al. Biodistribution and first clinical results of 18F-SiFAlin-TATE PET: A novel 18F-labeled somatostatin analog for imaging of neuroendocrine tumors. Eur. J. Nucl. Med. Mol. Imaging 2020, 47, 870–880. [Google Scholar] [CrossRef] [PubMed]
  156. Pauwels, E.; Cleeren, F.; Tshibangu, T.; Koole, M.; Serdons, K.; Dekervel, J.; Van Cutsem, E.; Verslype, C.; Van Laere, K.; Bormans, G.; et al. [18F]AlF-NOTA-octreotide PET imaging: Biodistribution, dosimetry and first comparison with [68Ga]Ga-DOTATATE in neuroendocrine tumour patients. Eur. J. Nucl. Med. Mol. Imaging 2020. [Google Scholar] [CrossRef]
  157. Kemerink, G.J.; Visser, M.G.; Franssen, R.; Beijer, E.; Zamburlini, M.; Halders, S.G.; Brans, B.; Mottaghy, F.M.; Teule, G.J. Effect of the positron range of 18F, 68Ga and 124I on PET/CT in lung-equivalent materials. Eur. J. Nucl. Med. Mol. Imaging 2011, 38, 940–948. [Google Scholar] [CrossRef]
  158. Uccelli, L.; Martini, P.; Cittanti, C.; Carnevale, A.; Missiroli, L.; Giganti, M.; Bartolomei, M.; Boschi, A. Therapeutic Radiometals: Worldwide Scientific Literature Trend Analysis (2008–2018). Molecules 2019, 24, 640. [Google Scholar] [CrossRef] [Green Version]
  159. Cremonesi, M.; Ferrari, M.E.; Bodei, L.; Chiesa, C.; Sarnelli, A.; Garibaldi, C.; Pacilio, M.; Strigari, L.; Summers, P.E.; Orecchia, R.; et al. Correlation of dose with toxicity and tumour response to 90Y- and 177Lu-PRRT provides the basis for optimization through individualized treatment planning. Eur. J. Nucl. Med. Mol. Imaging 2018, 45, 2426–2441. [Google Scholar] [CrossRef]
  160. Otte, A.; Mueller-Brand, J.; Dellas, S.; Nitzsche, E.U.; Herrmann, R.; Maecke, H.R. Yttrium-90 labeled somatostatin-analogue for cancer treatment. Lancet 1998, 351, 417–418. [Google Scholar] [CrossRef]
  161. Otte, A.; Herrmann, R.; Heppeler, A.; Behe, M.; Jermann, E.; Powell, P.; Maecke, H.R.; Muller, J. Yttrium-90 DOTATOC: First clinical results. Eur. J. Nucl. Med. 1999, 26, 1439–1447. [Google Scholar] [CrossRef] [PubMed]
  162. Vinjamuri, S.; Gilbert, T.M.; Banks, M.; McKane, G.; Maltby, P.; Poston, G.; Weissman, H.; Palmer, D.H.; Vora, J.; Pritchard, D.M.; et al. Peptide Receptor Radionuclide Therapy With 90Y-DOTATATE/90Y-DOTATOC in Patients with Progressive Metastatic Neuroendocrine Tumours: Assessment of Response, Survival and Toxicity. Br. J. Cancer 2013, 108, 1440–1448. [Google Scholar] [CrossRef] [PubMed]
  163. Virgolini, I.; Britton, K.; Buscombe, J.; Moncayo, R.; Paganelli, G.; Riva, P. In- and Y-DOTA-lanreotide: Results and implications of the MAURITIUS trial. Semin. Nucl. Med. 2002, 32, 148–155. [Google Scholar] [CrossRef] [PubMed]
  164. Bodei, L.; Cremonesi, M.; Grana, C.M.; Chinol, M.; Baio, S.M.; Severi, S.; Paganelli, G. Yttrium-labeled peptides for therapy of NET. Eur. J. Nucl. Med. Mol. Imaging 2012, 39, S93–S102. [Google Scholar] [CrossRef]
  165. Gabriel, M.; Nilica, B.; Kaiser, B.; Virgolini, I.J. Twelve-Year Follow-up After Peptide Receptor Radionuclide Therapy. J. Nucl. Med. 2019, 60, 524–529. [Google Scholar] [CrossRef] [Green Version]
  166. Baum, R.P.; Kluge, A.W.; Kulkarni, H.; Schorr-Neufing, U.; Niepsch, K.; Bitterlich, N.; van Echteld, C.J. [177Lu-DOTA]0-D-Phe1-Tyr3-Octreotide (177Lu-DOTATOC) For Peptide Receptor Radiotherapy in Patients with Advanced Neuroendocrine Tumours: A Phase-II Study. Theranostics 2016, 6, 501–510. [Google Scholar] [CrossRef]
  167. Esser, J.P.; Krenning, E.P.; Teunissen, J.J.; Kooij, P.P.; van Gameren, A.L.; Bakker, W.H.; Kwekkeboom, D.J. Comparison of [177Lu-DOTA0,Tyr3]octreotate and [177Lu-DOTA0,Tyr3]octreotide: Which peptide is preferable for PRRT? Eur. J. Nucl. Med. Mol. Imaging 2006, 33, 1346–1351. [Google Scholar] [CrossRef]
  168. de Jong, M.; Breeman, W.A.; Valkema, R.; Bernard, B.F.; Krenning, E.P. Combination radionuclide therapy using 177Lu- and 90Y-labeled somatostatin analogs. J. Nucl. Med. 2005, 46, 13S–17S. [Google Scholar]
  169. Kunikowska, J.; Królicki, L.; Hubalewska-Dydejczyk, A.; Mikołajczak, R.; Sowa-Staszczak, A.; Pawlak, D. Clinical results of radionuclide therapy of neuroendocrine tumours with 90Y-DOTATATE and tandem 90Y/177Lu-DOTATATE: Which is a better therapy option? Eur. J. Nucl. Med. Mol. Imaging 2011, 38, 1788–1797. [Google Scholar] [CrossRef] [Green Version]
  170. Kunikowska, J.; Zemczak, A.; Kołodziej, M.; Gut, P.; Łoń, I.; Pawlak, D.; Mikołajczak, R.; Kamiński, G.; Ruchała, M.; Kos-Kudła, B.; et al. Tandem peptide receptor radionuclide therapy using 90Y/177Lu-DOTATATE for neuroendocrine tumors efficacy and side-effects—Polish multicenter experience. Eur. J. Nucl. Med. Mol. Imaging 2020, 47, 922–933. [Google Scholar] [CrossRef] [Green Version]
  171. Brabander, T.; Nonnekens, J.; Hofland, J. The next generation of peptide receptor radionuclide therapy. Endocr. Relat. Cancer 2019, 26, C7–C11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Adant, S.; Shah, G.M.; Beauregard, J.M. Combination treatments to enhance peptide receptor radionuclide therapy of neuroendocrine tumours. Eur. J. Nucl. Med. Mol. Imaging 2020, 47, 907–921. [Google Scholar] [CrossRef] [PubMed]
  173. Strosberg, J.; El-Haddad, G.; Wolin, E.; Hendifar, A.; Yao, J.; Chasen, B.; Mittra, E.; Kunz, P.L.; Kulke, M.H.; Jacene, H.; et al. Phase 3 Trial of 177Lu-Dotatate for Midgut Neuroendocrine Tumors. N. Engl. J. Med. 2017, 376, 125–135. [Google Scholar] [CrossRef] [PubMed]
  174. Hennrich, U.; Kopka, K. Lutathera®: The First FDA- and EMA-Approved Radiopharmaceutical for Peptide Receptor Radionuclide Therapy. Pharmaceuticals 2019, 12, 114. [Google Scholar] [CrossRef] [Green Version]
  175. Werner, R.A.; Bluemel, C.; Allen-Auerbach, M.S.; Higuchi, T.; Herrmann, K. 68Gallium- and 90Yttrium-/177Lutetium: “theranostic twins” for diagnosis and treatment of NETs. Ann. Nucl. Med. 2015, 29, 1–7. [Google Scholar] [CrossRef] [Green Version]
  176. Waseem, N.; Aparici, C.M.; Kunz, P.L. Evaluating the Role of Theranostics in Grade 3 Neuroendocrine Neoplasms. J. Nucl. Med. 2019, 60, 882–891. [Google Scholar] [CrossRef] [Green Version]
  177. Sorbye, H.; Kong, G.; Grozinsky-Glasberg, S. PRRT in high-grade gastroenteropancreatic neuroendocrine neoplasms (WHO G3). Endocr. Relat. Cancer 2020, 27, R67–R77. [Google Scholar] [CrossRef]
  178. Mak, I.Y.F.; Hayes, A.R.; Khoo, B.; Grossman, A. Peptide Receptor Radionuclide Therapy as a Novel Treatment for Metastatic and Invasive Phaeochromocytoma and Paraganglioma. Neuroendocrinology 2019, 109, 287–298. [Google Scholar] [CrossRef]
  179. Vyakaranam, A.R.; Crona, J.; Norlén, O.; Granberg, D.; Garske-Román, U.; Sandström, M.; Fröss-Baron, K.; Thiis-Evensen, E.; Hellman, P.; Sundin, A. Favorable Outcome in Patients with Pheochromocytoma and Paraganglioma Treated with 177Lu-DOTATATE. Cancers 2019, 11, 909. [Google Scholar] [CrossRef] [Green Version]
  180. Lepareur, N.; Lacœuille, F.; Bouvry, C.; Hindré, F.; Garcion, E.; Chérel, M.; Noiret, N.; Garin, E.; Knapp, F.F.R., Jr. Rhenium-188 Labeled Radiopharmaceuticals: Current Clinical Applications in Oncology and Promising Perspectives. Front. Med. 2019, 6, 132. [Google Scholar] [CrossRef] [Green Version]
  181. Zamora, P.O.; Gulhke, S.; Bender, H.; Diekmann, D.; Rhodes, B.A.; Biersack, H.J.; Knapp, F.F., Jr. Experimental radiotherapy of receptor-positive human prostate adenocarcinoma with 188Re-RC-160, a directly-radiolabeled somatostatin analogue. Int. J. Cancer 1996, 65, 214–220. [Google Scholar] [CrossRef]
  182. Zamora, P.O.; Bender, H.; Gulhke, S.; Marek, M.J.; Knapp, F.F., Jr.; Rhodes, B.A.; Biersack, H.J. Pre-clinical experience with Re-188-RC-160, a radiolabeled somatostatin analog for use in peptide-targeted radiotherapy. Anticancer Res. 1997, 17, 1803–1808. [Google Scholar] [PubMed]
  183. Arteaga de Murphy, C.; Pedraza-López, M.; Ferro-Flores, G.; Murphy-Stack, E.; Chávez-Mercado, L.; Ascencio, J.A.; García-Salinas, L.; Hernández-Gutiérrez, S. Uptake of 188Re-beta-naphthyl-peptide in cervical carcinoma tumours in athymic mice. Nucl. Med. Biol. 2001, 28, 319–326. [Google Scholar] [CrossRef]
  184. Molina-Trinidad, E.M.; de Murphy, C.A.; Ferro-Flores, G.; Murphy-Stack, E.; Jung-Cook, H. Radiopharmacokinetic and dosimetric parameters of 188Re-lanreotide in athymic mice with induced human cancer tumors. Int. J. Pharm. 2006, 310, 125–130. [Google Scholar] [CrossRef]
  185. Molina-Trinidad, E.M.; de Murphy, C.A.; Jung-Cook, H.; Stack, E.M.; Pedraza-Lopez, M.; Morales-Marquez, J.L.; Serrano, G.V. Therapeutic 188Re-lanreotide: Determination of radiopharmacokinetic parameters in rats. J. Pharm. Pharmacol. 2010, 62, 456–461. [Google Scholar] [CrossRef]
  186. Cyr, J.E.; Pearson, D.A.; Wilson, D.M.; Nelson, C.A.; Guaraldi, M.; Azure, M.T.; Lister-James, J.; Dinkelborg, L.M.; Dean, R.T. Somatostatin receptor-binding peptides suitable for tumour radiotherapy with Re-188 or Re-186. Chemistry and initial biological studies. J. Med. Chem. 2007, 50, 1354–1364. [Google Scholar] [CrossRef]
  187. Edelman, M.J.; Clamon, G.; Kahn, D.; Magram, M.; Lister-James, J.; Line, B.R. Targeted radiopharmaceutical therapy for advanced lung cancer: Phase I trial of rhenium Re188 P2045, a somatostatin analog. J. Thorac. Oncol. 2009, 4, 1550–1554. [Google Scholar] [CrossRef] [Green Version]
  188. Nelson, C.A.; Azure, M.T.; Adams, C.T.; Zinn, K.R. The somatostatin analog 188Re-P2045 inhibits the growth of AR42J pancreatic tumor xenografts. J. Nucl. Med. 2014, 55, 2020–2025. [Google Scholar] [CrossRef] [Green Version]
  189. Champion, C.; Quinto, M.A.; Morgat, C.; Zanotti-Fregonara, P.; Hindié, E. Comparison between Three Promising ß-emitting Radionuclides, 67Cu, 47Sc and 161Tb, with Emphasis on Doses Delivered to Minimal Residual Disease. Theranostics 2016, 6, 1611–1618. [Google Scholar] [CrossRef] [Green Version]
  190. De Jong, M.; Breeman, W.A.; Bernard, B.F.; Rolleman, E.J.; Hofland, L.J.; Visser, T.J.; Setyono-Han, B.; Bakker, W.H.; van der Pluijm, M.E.; Krenning, E.P. Evaluation in vitro and in rats of 161Tb-DTPA-octreotide, a somatostatin analogue with potential for intraoperative scanning and radiotherapy. Eur. J. Nucl. Med. 1995, 22, 608–616. [Google Scholar] [CrossRef] [Green Version]
  191. Loveless, C.S.; Radford, L.L.; Ferran, S.J.; Queern, S.L.; Shepherd, M.R.; Lapi, S.E. Photonuclear production, chemistry, and in vitro evaluation of the theranostic radionuclide 47Sc. EJNMMI Res. 2019, 9, 42. [Google Scholar] [CrossRef] [PubMed]
  192. Tafreshi, N.K.; Doligalski, M.L.; Tichacek, C.J.; Pandya, D.N.; Budzevich, M.M.; El-Haddad, G.; Khushalani, N.I.; Moros, E.G.; McLaughlin, M.L.; Wadas, T.J.; et al. Development of Targeted Alpha Particle Therapy for Solid Tumors. Molecules 2019, 24, 4314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Nayak, T.; Norenberg, J.; Anderson, T.; Atcher, R. A comparison of high- versus low-linear energy transfer somatostatin receptor targeted radionuclide therapy in vitro. Cancer Biother. Radiopharm. 2005, 20, 52–57. [Google Scholar] [CrossRef] [PubMed]
  194. Nayak, T.K.; Norenberg, J.P.; Anderson, T.L.; Prossnitz, E.R.; Stabin, M.G.; Atcher, R.W. Somatostatin-receptor-targeted α-emitting 213Bi is therapeutically more effective than β--emitting 177Lu in human pancreatic adenocarcinoma cells. Nucl. Med. Biol. 2007, 34, 185–193. [Google Scholar] [CrossRef] [PubMed]
  195. Chan, H.S.; de Blois, E.; Morgenstern, A.; Bruchertseifer, F.; de Jong, M.; Breeman, W.; Konijnenberg, M. In Vitro comparison of 213Bi- and 177Lu-radiation for peptide receptor radionuclide therapy. PLoS ONE 2017, 12, e0181473. [Google Scholar] [CrossRef]
  196. Norenberg, J.P.; Krenning, B.J.; Konings, I.R.H.M.; Kusewitt, D.F.; Nayak, T.K.; Anderson, T.L.; de Jong, M.; Garmestani, K.; Brechbiel, M.W.; Kvols, L.K. 213Bi-[DOTA0, Tyr3]Octreotide Peptide Receptor Radionuclide Therapy of Pancreatic Tumors in a Preclinical Animal Model. Clin. Cancer Res. 2006, 12, 897–903. [Google Scholar] [CrossRef] [Green Version]
  197. Miederer, M.; Henriksen, G.; Alke, A.; Mossbrugger, I.; Quintanilla-Martinez, L.; Senekowitsch-Schmidtke, R.; Essler, M. Preclinical evaluation of the alpha-particle generator nuclide 225Ac for somatostatin receptor radiotherapy of neuroendocrine tumors. Clin. Cancer Res. 2008, 14, 3555–3561. [Google Scholar] [CrossRef] [Green Version]
  198. Chan, H.S.; Konijnenberg, M.W.; de Blois, E.; Koelewijn, S.; Baum, R.P.; Morgenstern, A.; Bruchertseifer, F.; Breeman, W.A.; de Jong, M. Influence of tumour size on the efficacy of targeted alpha therapy with 213Bi-[DOTA0,Tyr3]-octreotate. EJNMMI Res. 2016, 6, 6. [Google Scholar] [CrossRef]
  199. Kratochwil, C.; Giesel, F.L.; Bruchertseifer, F.; Mier, W.; Apostolidis, C.; Boll, R.; Murphy, K.; Haberkorn, U.; Morgenstern, A. 213Bi-DOTATOC receptor-targeted alpha-radionuclide therapy induces remission in neuroendocrine tumours refractory to beta radiation: A first-in-human experience. Eur. J. Nucl. Med. Mol. Imaging 2014, 41, 2106–2119. [Google Scholar] [CrossRef] [Green Version]
  200. Zhang, J.; Singh, A.; Kulkarni, H.R.; Schuchardt, C.; Müller, D.; Wester, H.J.; Maina, T.; Rösch, F.; van der Meulen, N.P.; Müller, C.; et al. From Bench to Bedside-The Bad Berka Experience with First-in-Human Studies. Semin. Nucl. Med. 2019, 49, 422–437. [Google Scholar] [CrossRef]
  201. Ballal, S.; Yadav, M.P.; Bal, C.; Sahoo, R.K.; Tripathi, M. Broadening horizons with 225Ac-DOTATATE targeted alpha therapy for gastroenteropancreatic neuroendocrine tumour patients stable or refractory to 177Lu-DOTATATE PRRT: First clinical experience on the efficacy and safety. Eur. J. Nucl. Med. Mol. Imaging 2020, 47, 934–946. [Google Scholar] [CrossRef] [PubMed]
  202. Stallons, T.A.R.; Saidi, A.; Tworowska, I.; Delpassand, E.S.; Torgue, J.J. Preclinical Investigation of 212Pb-DOTAMTATE for Peptide Receptor Radionuclide Therapy in a Neuroendocrine Tumor Model. Mol. Cancer Ther. 2019, 18, 1012–1021. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Tworowska, I.; Delpassand, E.S.; Bolek, L.; Shanoon, F.; Sgouros, G.; Frey, E.; He, B.; Muzammil, A.; Ghaly, M.; Stallons, T.; et al. Targeted Alpha-emitter Therapy of Neuroendocrine Tumors using 212Pb-octreotate (AlphaMedix™). J. Med. Imaging Radiat. Sci. 2019, 50, S34. [Google Scholar] [CrossRef]
  204. Guérard, F.; Gestin, J.F.; Brechbiel, M.W. Production of [211At]-astatinated radiopharmaceuticals and applications in targeted α-particle therapy. Cancer Biother. Radiopharm. 2013, 28, 1–20. [Google Scholar] [CrossRef] [Green Version]
  205. Vaidyanathan, G.; Boskovitz, A.; Shankar, S.; Zalutsky, M.R. Radioiodine and 211At-labeled guanidinomethyl halobenzoyl octreotate conjugates: Potential peptide radiotherapeutics for somatostatin receptor-positive cancers. Peptides 2004, 25, 2087–2097. [Google Scholar] [CrossRef] [PubMed]
  206. Vaidyanathan, G.; Affleck, D.J.; Schottelius, M.; Wester, H.; Friedman, H.S.; Zalutsky, M.R. Synthesis and evaluation of glycosylated octreotate analogues labeled with radioiodine and 211At via a tin precursor. Bioconjugate Chem. 2006, 17, 195–203. [Google Scholar] [CrossRef]
  207. Zhao, B.; Qin, S.; Chai, L.; Lu, G.; Yang, Y.; Cai, H.; Yuan, X.; Fan, S.; Huang, Q.; Yu, F. Evaluation of astatine-211-labeled octreotide as a potential radiotherapeutic agent for NSCLC treatment. Bioorg. Med. Chem. 2018, 26, 1086–1091. [Google Scholar] [CrossRef]
  208. Krenning, E.P.; Valkema, R.; Kooij, P.P.; Breeman, W.A.; Bakker, W.H.; de Herder, W.W.; van Eijck, C.H.; Kwekkeboom, D.J.; de Jong, M.; Pauwels, S. Scintigraphy and radionuclide therapy with [indium-111-labeled-diethyltriamine penta-acetic acid-D-Phe1]-octreotide. Ital. J. Gastroenterol. Hepatol. 1999, 31, S219–S223. [Google Scholar]
  209. Valkema, R.; De Jong, M.; Bakker, W.H.; Breeman, W.A.; Kooij, P.P.; Lugtenburg, P.J.; De Jong, F.H.; Christiansen, A.; Kam, B.L.; De Herder, W.W.; et al. Phase I study of peptide receptor radionuclide therapy with [In-DTPA]octreotide: The Rotterdam experience. Semin. Nucl. Med. 2002, 32, 110–122. [Google Scholar] [CrossRef]
  210. Anthony, L.B.; Woltering, E.A.; Espenan, G.D.; Cronin, M.D.; Maloney, T.J.; McCarthy, K.E. Indium-111-pentetreotide prolongs survival in gastroenteropancreatic malignancies. Semin. Nucl. Med. 2002, 32, 123–132. [Google Scholar] [CrossRef]
  211. Buscombe, J.R.; Caplin, M.E.; Hilson, A.J. Long-term efficacy of high-activity 111In-pentetreotide therapy in patients with disseminated neuroendocrine tumors. J. Nucl. Med. 2003, 44, 1–6. [Google Scholar] [PubMed]
  212. Lewington, V.J. Targeted radionuclide therapy for neuroendocrine tumours. Endocr. Relat. Cancer 2003, 10, 497–501. [Google Scholar] [CrossRef] [PubMed]
  213. Capello, A.; Krenning, E.; Bernard, B.; Reubi, J.C.; Breeman, W.; de Jong, M. 111In-labeled somatostatin analogues in a rat tumour model: Somatostatin receptor status and effects of peptide receptor radionuclide therapy. Eur. J. Nucl. Med. Mol. Imaging 2005, 32, 1288–1295. [Google Scholar] [CrossRef]
  214. 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] [PubMed] [Green Version]
  215. Waser, B.; Tamma, M.L.; Cescato, R.; Maecke, H.R.; Reubi, J.C. Highly efficient in vivo agonist induced internalization of sst2 receptors in somatostatin target tissues. J. Nucl. Med. 2009, 50, 936941. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Cescato, R.; Schulz, S.; Waser, B.; Eltschinger, V.; Rivier, J.E.; Wester, H.J.; Culler, M.; Ginj, M.; Liu, Q.; Schonbrunn, A.; et al. Internalization of sst2, sst3, and sst5 receptors: Effects of somatostatin agonists and antagonists. J. Nucl. Med. 2006, 47, 502–511. [Google Scholar] [PubMed]
  217. Fani, M.; Peitl, P.K.; Velikyan, I. Current status of radiopharmaceuticals for the theranostics of neuroendocrine neoplasms. Pharmaceuticals 2017, 10, 30. [Google Scholar] [CrossRef] [PubMed]
  218. Reubi, J.C.; Waser, B.; Mäcke, H.; Rivier, J. Highly Increased 125I-JR11 Antagonist Binding In Vitro Reveals Novel Indications for sst2 Targeting in Human Cancers. J. Nucl. Med. 2017, 58, 300–306. [Google Scholar] [CrossRef] [Green Version]
  219. Dude, I.; Zhang, Z.; Rousseau, J.; Hundal-Jabal, N.; Colpo, N.; Merkens, H.; Lin, K.S.; Bénard, F. Evaluation of agonist and antagonist radioligands for somatostatin receptor imaging of breast cancer using positron emission tomography. EJNMMI Radiopharm. Chem. 2017, 2, 4. [Google Scholar] [CrossRef]
  220. Rylova, S.N.; Stoykow, C.; Del Pozzo, L.; Abiraj, K.; Tamma, M.L.; Kiefer, Y.; Fani, M.; Maecke, H.R. The somatostatin receptor 2 antagonist 64Cu-NODAGA-JR11 outperforms 64Cu-DOTA-TATE in a mouse xenograft model. PLoS ONE 2018, 13, e0195802. [Google Scholar] [CrossRef] [Green Version]
  221. Krebs, S.; Pandit-Taskar, N.; Reidy, D.; Beattie, B.J.; Lyashchenko, S.K.; Lewis, J.S.; Bodei, L.; Weber, W.A.; O’Donoghue, J.A. Biodistribution and radiation dose estimates for 68Ga-DOTA-JR11 in patients with metastatic neuroendocrine tumors. Eur. J. Nucl. Med. Mol. Imaging 2019, 46, 677–685. [Google Scholar] [CrossRef] [PubMed]
  222. 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] [PubMed] [Green Version]
  223. Fani, M.; Del Pozzo, L.; Abiraj, K.; Mansi, R.; Tamma, M.L.; Cescato, R.; Waser, B.; Weber, W.A.; Reubi, J.C.; Maecke, H.R. PET of somatostatin receptor-positive tumors using 64Cu- and 68Ga-somatostatin antagonists: The chelate makes the difference. J. Nucl. Med. 2011, 52, 1110–1118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. 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] [Green Version]
  225. Reubi, J.C.; Erchegyi, J.; Cescato, R.; Waser, B.; Rivier, J.E. Switch from antagonist to agonist after addition of a DOTA chelator to a somatostatin analog. Eur. J. Nucl. Med. Mol. Imaging 2010, 37, 15511558. [Google Scholar] [CrossRef] [Green Version]
  226. Radford, L.; Gallazzi, F.; Watkinson, L.; Carmack, T.; Berendzen, A.; Lewis, M.R.; Jurisson, S.S.; Papagiannopoulou, D.; Hennkens, H.M. Synthesis and evaluation of a 99mTc tricarbonyl-labeled somatostatin receptor-targeting antagonist peptide for imaging of neuroendocrine tumors. Nucl. Med. Biol. 2017, 47, 4–9. [Google Scholar] [CrossRef] [Green Version]
  227. Wild, D.; Fani, M.; Fischer, R.; Del Pozzo, L.; Kaul, F.; Krebs, S.; Fischer, R.; Rivier, J.E.; Reubi, J.C.; Maecke, H.R.; et al. Comparison of somatostatin receptor agonist and antagonist for peptide receptor radionuclide therapy: A pilot study. J. Nucl. Med. 2014, 55, 1248–1252. [Google Scholar] [CrossRef] [Green Version]
  228. Fani, M.; Nicolas, G.P.; Wild, D. Somatostatin Receptor Antagonists for Imaging and Therapy. J. Nucl. Med. 2017, 58, 61S–66S. [Google Scholar] [CrossRef]
  229. Mansi, R.; Fani, M. Design and development of the theranostic pair 177Lu-OPS201/68Ga-OPS202 for targeting somatostatin receptor expressing tumors. J. Label. Compd. Radiopharm. 2019, 62, 635–645. [Google Scholar] [CrossRef]
  230. Nicolas, G.P.; Beykan, S.; Bouterfa, H.; Kaufmann, J.; Bauman, A.; Lassmann, M.; Reubi, J.C.; Rivier, J.E.F.; Maecke, H.R.; Fani, M.; et al. Safety, Biodistribution, and Radiation Dosimetry of 68Ga-OPS202 in Patients with Gastroenteropancreatic Neuroendocrine Tumors: A Prospective Phase I Imaging Study. J. Nucl. Med. 2018, 59, 909–914. [Google Scholar] [CrossRef] [Green Version]
  231. Nicolas, G.P.; Schreiter, N.; Kaul, F.; Uiters, J.; Bouterfa, H.; Kaufmann, J.; Erlanger, T.E.; Cathomas, R.; Christ, E.; Fani, M.; et al. Sensitivity Comparison of 68Ga-OPS202 and 68Ga-DOTATOC PET/CT in Patients with Gastroenteropancreatic Neuroendocrine Tumors: A Prospective Phase II Imaging Study. J. Nucl. Med. 2018, 59, 915–921. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Reidy-Lagunes, D.; Pandit-Taskar, N.; O’Donoghue, J.A.; Krebs, S.; Staton, K.D.; Lyashchenko, S.K.; Lewis, J.S.; Raj, N.; Gönen, M.; Lohrmann, C.; et al. Phase I Trial of Well-Differentiated Neuroendocrine Tumors (NETs) with Radiolabeled Somatostatin Antagonist 177Lu-Satoreotide Tetraxetan. Clin. Cancer Res. 2019, 25, 6939–6947. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Rangger, C.; Haubner, R. Radiolabeled Peptides for Positron Emission Tomography and Endoradiotherapy in Oncology. Pharmaceuticals 2020, 13, 22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Villard, L.; Romer, A.; Marincek, N.; Brunner, P.; Koller, M.T.; Schindler, C.; Ng, Q.K.T.; Mäcke, H.R.; Müller-Brand, J.; Rochlitz, C.; et al. Cohort study of somatostatin-based radiopeptide therapy with [90Y-DOTA]-TOC versus [90Y-DOTA]-TOC plus [177Lu-DOTA]-TOC in neuroendocrine cancers. J. Clin. Oncol. 2012, 30, 1100–1106. [Google Scholar] [CrossRef]
  235. Gill, M.R.; Falzone, N.; Du, Y.; Vallis, K.A. Targeted radionuclide therapy in combined-modality regimens. Lancet Oncol. 2017, 18, e414–e423. [Google Scholar] [CrossRef]
  236. Hartrampf, P.E.; Hänscheid, H.; Kertels, O.; Schirbel, A.; Kreissl, M.C.; Flentje, M.; Sweeney, R.A.; Buck, A.K.; Polat, B.; Lapa, C. Long-term results of multimodal peptide receptor radionuclide therapy and fractionated external beam radiotherapy for treatment of advanced symptomatic meningioma. Clin. Transl. Radiat. Oncol. 2020, 22, 29–32. [Google Scholar] [CrossRef] [Green Version]
  237. Reubi, J.C.; Maecke, H.R. Approaches to multireceptor targeting: Hybrid radioligands, radioligand cocktails, and sequential radioligand applications. J. Nucl. Med. 2017, 58, 10S–16S. [Google Scholar] [CrossRef] [Green Version]
  238. Ghosh, S.C.; Rodriguez, M.; Carmon, K.S.; Voss, J.; Wilganowski, N.L.; Schonbrunn, A.; Azhdarinia, A. A Modular Dual-Labeling Scaffold That Retains Agonistic Properties for Somatostatin Receptor Targeting. J. Nucl. Med. 2017, 58, 1858–1864. [Google Scholar] [CrossRef] [Green Version]
  239. Langbein, T.; Weber, W.A.; Eiber, M. Future of Theranostics: An Outlook on Precision Oncology in Nuclear Medicine. J. Nucl. Med. 2019, 60, 13S–19S. [Google Scholar] [CrossRef] [Green Version]
  240. Jones, W.; Griffiths, K.; Barata, P.C.; Paller, C.J. PSMA Theranostics: Review of the Current Status of PSMA-Targeted Imaging and Radioligand Therapy. Cancers 2020, 12, 1367. [Google Scholar] [CrossRef]
Figure 1. Somatostatin receptors (SSTRs) biodistribution in the body (from The Human Protein Atlas https://www.proteinatlas.org/).
Figure 1. Somatostatin receptors (SSTRs) biodistribution in the body (from The Human Protein Atlas https://www.proteinatlas.org/).
Molecules 25 04012 g001
Figure 2. Schematic representation of the signaling pathways induced by somatostatin receptors activation. Green arrows: activated pathways; red arrows: inhibited pathways. Adapted from [8].
Figure 2. Schematic representation of the signaling pathways induced by somatostatin receptors activation. Green arrows: activated pathways; red arrows: inhibited pathways. Adapted from [8].
Molecules 25 04012 g002
Figure 3. Chemical structures of SRIF-14, SRIF-28, and selected examples of somatostatin analogs.
Figure 3. Chemical structures of SRIF-14, SRIF-28, and selected examples of somatostatin analogs.
Molecules 25 04012 g003
Figure 4. Schematic design of a radiometallated bioconjugate.
Figure 4. Schematic design of a radiometallated bioconjugate.
Molecules 25 04012 g004
Figure 5. Representative (but not exhaustive) examples of acyclic and macrocyclic polyamino and polyaminocarboxylic chelator families and their derivatives.
Figure 5. Representative (but not exhaustive) examples of acyclic and macrocyclic polyamino and polyaminocarboxylic chelator families and their derivatives.
Molecules 25 04012 g005
Figure 6. Structure of [111In]-pentetreotide (Octreoscan®).
Figure 6. Structure of [111In]-pentetreotide (Octreoscan®).
Molecules 25 04012 g006
Figure 7. Structures of the three main systems radiolabeled with gallium-68.
Figure 7. Structures of the three main systems radiolabeled with gallium-68.
Molecules 25 04012 g007
Figure 8. [99mTc]-labeled somatostatin analogs.
Figure 8. [99mTc]-labeled somatostatin analogs.
Molecules 25 04012 g008
Figure 9. [18F]-labeled somatostatin analogs.
Figure 9. [18F]-labeled somatostatin analogs.
Molecules 25 04012 g009
Figure 10. (A) [68Ga]-DOTATOC (Somakit®) and (B) [177Lu]-DOTATATE (Lutathera®, cures 1, 2, and 3) imaging of a patient treated for progressive metastatic midgut NET (images courtesy of Centre Eugene Marquis, Rennes, France).
Figure 10. (A) [68Ga]-DOTATOC (Somakit®) and (B) [177Lu]-DOTATATE (Lutathera®, cures 1, 2, and 3) imaging of a patient treated for progressive metastatic midgut NET (images courtesy of Centre Eugene Marquis, Rennes, France).
Molecules 25 04012 g010
Figure 11. [188Re]-P2045.
Figure 11. [188Re]-P2045.
Molecules 25 04012 g011
Figure 12. [212Pb]-DOTAMTATE.
Figure 12. [212Pb]-DOTAMTATE.
Molecules 25 04012 g012
Figure 13. Comparison between [68Ga]-OPS202 (A,B) and [68Ga]-DOTATOC (C,D) PET/CT images of the same patient with ileal neuroendocrine tumours, showing bilobar liver metastases (from Rangger et al. [233]).
Figure 13. Comparison between [68Ga]-OPS202 (A,B) and [68Ga]-DOTATOC (C,D) PET/CT images of the same patient with ileal neuroendocrine tumours, showing bilobar liver metastases (from Rangger et al. [233]).
Molecules 25 04012 g013
Table 1. SSTRs expression in different tumor types.
Table 1. SSTRs expression in different tumor types.
Tumor TypeSSTR ExpressionRef
Astrocytoma+[17]
Breast carcinoma+ (SSTR2)[11]
Cholangiocarcinoma+ (SSTR2)[18]
Colorectal carcinoma-[17]
Endometrial carcinoma-[17]
Ependymoma+ (SSTR1, SSTR5)[11]
Esophageal carcinoma-[17]
Ewing sarcoma-[17]
Exocrine pancreatic tumor-[17]
Gastric carcinoma+ (SSTR1 > SSTR2, SSTR5)[11]
Gastrinoma+ (SSTR2)[17]
Glioblastoma-[17]
Growth hormone-producing pituitary adenoma+ (SSTR2, SSTR5)[17]
Gut carcinoid+ (SSTR2 > SSTR1, SSTR5)[17]
Hepatocellular carcinoma+ (SSTR2, SSTR5)[19]
Insulinoma+ (SSTR1, SSTR2, SSTR3)[20]
Leiomyoma+[17]
Lymphoma+ (SSTR2)[11]
Medullary thyroid carcinoma+ (SSTR2)[11]
Medulloblastoma+ (SSTR2)[17]
Meningioma+ (SSTR2)[17]
Neuroblastoma+ (SSTR2)[17]
Non-functioning pituitary adenoma+ (SSTR3 > SSTR2)[17]
Non-small cell lung cancer-[17]
Ovarian carcinoma+[17]
Paraganglioma+ (SSTR2)[17]
Pheochromocytoma+ (SSTR1, SSTR2)[17]
Prostate carcinoma+ (SSTR1)[17]
Renal cell carcinoma+ (SSTR2)[11]
Small cell lung cancer+ (SSTR2)[17]
Urinary bladder carcinoma-[17]
Bold +, receptors with particularly high density and incidence. Subtypes preferentially expressed are listed in parentheses, only when compelling evidence is available (immunohistochemistry or autoradiography). Adapted from [11] and [17].
Table 2. Some of the main radionuclides studied for imaging and therapy (SPECT—Single-Photon Emission Computed Tomography; PET—Positron Emission Tomography).
Table 2. Some of the main radionuclides studied for imaging and therapy (SPECT—Single-Photon Emission Computed Tomography; PET—Positron Emission Tomography).
RadionuclideHalf-Life (h)Type of EmissionEnergy of Emitted Radiation (keV)SourceApplication
99mTc6.01γ140GeneratorSPECT imaging
111In67.4γ172, 245CyclotronSPECT imaging
18F1.83β+634CyclotronPET imaging
64Cu12.7β+/γ/β-653CyclotronPET imaging
68Ga1.1β+1190Generator/CyclotronPET imaging
90Y64.1β2284GeneratorTherapy
177Lu160.8β497CyclotronTherapy
188Re17β2118GeneratorTherapy
211At7.2α5870CyclotronTherapy
225Ac238α5830GeneratorTherapy
Table 3. Peptidic sequences of the main somatostatin agonist analogs. Differences towards Octreotide (OC) are highlighted in red.
Table 3. Peptidic sequences of the main somatostatin agonist analogs. Differences towards Octreotide (OC) are highlighted in red.
PeptidePeptidic Sequence
OC
Octreotide
d-Phe1-cyclo(Cys2-Phe3-d-Trp4-Lys5-Thr6-Cys7)Thr(ol)8
LAN
Lanreotide
β-d-Nal1-cyclo(Cys2-Tyr3-d-Trp4-Lys5-Val6-Cys7)Thr8-NH2
VAP
Vapreotide
d-Phe1-cyclo(Cys2-Phe3-d-Trp4-Lys5-Val6-Cys7)Trp8-NH2
TOC
[Tyr3]-Octreotide
d-Phe1-cyclo(Cys2-Tyr3-d-Trp4-Lys5-Thr6-Cys7)Thr(ol)8
TATE
[Tyr3]-Octreotate
d-Phe1-cyclo(Cys2-Tyr3-d-Trp4-Lys5-Thr6-Cys7)Thr8
NOC
[1-Nal3]-Octreotide
d-Phe1-cyclo(Cys2-1-Nal3-d-Trp4-Lys5-Thr6-Cys7)Thr(ol)8
NOC-ATE
[1-Nal3, Thr8]-Octreotide
d-Phe1-cyclo(Cys2-1-Nal3-d-Trp4-Lys5-Thr6-Cys7)Thr8
BOC
[BzThi3]-Octreotide
d-Phe1-cyclo(Cys2-BzThi3-d-Trp4-Lys5-Thr6-Cys7)Thr(ol)8
BOC-ATE
[BzThi3, Thr8]-Octreotide
d-Phe1-cyclo(Cys2-BzThi3-d-Trp4-Lys5-Thr6-Cys7)Thr8
Table 4. Main somatostatin antagonist analogs. Differences towards octreotide (OC) are highlighted in red.
Table 4. Main somatostatin antagonist analogs. Differences towards octreotide (OC) are highlighted in red.
Antagonist PeptidePeptidic Sequence
Sst2-ANT (BASS)p-NO2-Phe1-cyclo(d-Cys2-Tyr3-d-Trp4-Lys5-Thr6-Cys7)d-Tyr8-NH2
LM3p-Cl-Phe1-cyclo(d-Cys2-Tyr3-d-Aph4(Cbm)-Lys5-Thr6-Cys7)d-Tyr8-NH2
JR10p-NO2-Phe1-cyclo(d-Cys2-Tyr3-d-Aph4(Cbm)-Lys5-Thr6-Cys7)d-Tyr8-NH2
JR11 (Satoreotide)p-Cl-Phe1-cyclo(d-Cys2-Aph3(Hor)-d-Aph4(Cbm)-Lys5-Thr6-Cys7)d-Tyr8-NH2

Share and Cite

MDPI and ACS Style

Eychenne, R.; Bouvry, C.; Bourgeois, M.; Loyer, P.; Benoist, E.; Lepareur, N. Overview of Radiolabeled Somatostatin Analogs for Cancer Imaging and Therapy. Molecules 2020, 25, 4012. https://doi.org/10.3390/molecules25174012

AMA Style

Eychenne R, Bouvry C, Bourgeois M, Loyer P, Benoist E, Lepareur N. Overview of Radiolabeled Somatostatin Analogs for Cancer Imaging and Therapy. Molecules. 2020; 25(17):4012. https://doi.org/10.3390/molecules25174012

Chicago/Turabian Style

Eychenne, Romain, Christelle Bouvry, Mickael Bourgeois, Pascal Loyer, Eric Benoist, and Nicolas Lepareur. 2020. "Overview of Radiolabeled Somatostatin Analogs for Cancer Imaging and Therapy" Molecules 25, no. 17: 4012. https://doi.org/10.3390/molecules25174012

Article Metrics

Back to TopTop