Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 3716 2023-04-29 09:10:51 |
2 Reference format revised. Meta information modification 3716 2023-05-03 04:30:24 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Chakraborty, K.; Mondal, J.; An, J.M.; Park, J.; Lee, Y. Radionuclides and Radiolabelled Peptides for Cancer Therapeutics. Encyclopedia. Available online: https://encyclopedia.pub/entry/43646 (accessed on 19 August 2024).
Chakraborty K, Mondal J, An JM, Park J, Lee Y. Radionuclides and Radiolabelled Peptides for Cancer Therapeutics. Encyclopedia. Available at: https://encyclopedia.pub/entry/43646. Accessed August 19, 2024.
Chakraborty, Kushal, Jagannath Mondal, Jeong Man An, Jooho Park, Yong-Kyu Lee. "Radionuclides and Radiolabelled Peptides for Cancer Therapeutics" Encyclopedia, https://encyclopedia.pub/entry/43646 (accessed August 19, 2024).
Chakraborty, K., Mondal, J., An, J.M., Park, J., & Lee, Y. (2023, April 29). Radionuclides and Radiolabelled Peptides for Cancer Therapeutics. In Encyclopedia. https://encyclopedia.pub/entry/43646
Chakraborty, Kushal, et al. "Radionuclides and Radiolabelled Peptides for Cancer Therapeutics." Encyclopedia. Web. 29 April, 2023.
Radionuclides and Radiolabelled Peptides for Cancer Therapeutics
Edit

Radiopharmaceutical therapy, which can detect and treat tumours simultaneously, was introduced more than 80 years ago, and it has changed medical strategies with respect to cancer. Many radioactive radionuclides have been developed, and functional, molecularly modified radiolabelled peptides have been used to produce biomolecules and therapeutics that are vastly utilised in the field of radio medicine. Advanced technologies, such as conjugation of functional peptides or incorporation of radionuclides into chelating ligands, have been developed for advanced radiopharmaceutical cancer therapy. New radiolabelled conjugates for targeted radiotherapy have been designed to deliver radiation directly to cancer cells with improved specificity and minimal damage to the surrounding normal tissue. The development of new theragnostic radionuclides, which can be used for both imaging and therapy purposes, allows for more precise targeting and monitoring of the treatment response. The increased use of peptide receptor radionuclide therapy (PRRT) is also important in the targeting of specific receptors which are overexpressed in cancer cells.

anticancer therapy Radionuclides PRRT

1. Introduction

Because cancer is one of the diseases with the highest mortality rates, the battle against cancer has attracted the attention of many scientists. Researchers have investigated a wide variety of approaches to treat malignant tumours, including photothermal, photodynamic, sonodynamic, and immune therapy [1][2][3][4][5][6][7][8][9][10][11]. Despite the significant developments made in the fight against cancer in recent years, no single therapy has been successful in treating the disease entirely. Better patient selection, prediction of treatment response and tissue toxicity, and response evaluation are all possible through theragnostics, since diagnostic and therapeutic methods relating to the same precise molecular targets can be coupled. Although the term theragnostics was originally put forth back in 2002, the notion underlying theragnostics has been used and explored over the years [12]. Nuclear theragnostics involves the use of radioactive compounds for imaging biological phenomena by detecting the expression of disease-related targets such as cell surface receptors or membrane transporters, followed by the application of agents designed to deliver ionising radiation to the tissues that express these targets. For the purpose of detection, a positron-emitting nuclide is injected intravenously, and the number of positrons emitted in the body is monitored using a detection camera. Images of nuclide distribution provide insight into the illness. The nuclides vary in their physical characteristics [13]. For effective diagnosis and therapy, it is preferable to employ the nuclides most suited to the particular disease in question. This is why it is important to think about aspects such as the linker, dosage, medicinal drug to be conjugated with, etc., as well as the nuclides themselves. The same mechanism of action underlies all forms of nuclear medicine therapy in which radiopharmaceutical substances are employed to selectively target cancerous tissue, and radiation is conveyed to individual cells via chemical and/or biological adherence.
There are two main types of particle radiation that are used for medical purposes, namely, radiation with α and β particles, and both of them have a similar feature: destruction of malignant cells as a result of severe DNA detrition. Some isotopes (e.g., 131I, 186Re, 153Sm etc.) can emit both therapeutic β particles as well as γ rays, making simultaneous diagnosis and therapy possible [14][15][16][17][18][19]. Diagnostic and therapeutic radiopharmaceuticals are referred to as theragnostic pairs when they access the same cellular structure and biological process. 
Radionuclides (radioactive nuclides or radioisotopes) have been studied for the treatment and diagnosis of disease for over 100 years. However, the utilisation of radionuclides in pharmaceutical medicine has faced several problems and limitations, such as the insufficiency of their targeting mechanisms. Iodine-131 is a radionuclide that was used to treat thyroid cancer in 1946, which was a great breakthrough in pharmaceutical medicine [20]. Other attempts have been made to develop radionuclides, such as iodine-131, but they were not very successful. Later, advancements in pharmaceutical medicine improved the use of radionuclides. The newly developed selective targeting delivery of radionuclides can prevent the occurrence of unfavourable delivery, thus reducing their side effects [21]. Moreover, the specific targeting delivery of radionuclides enhances the imaging quality at sites of interest, such as that of the tumour [22][23][24][25][26][27][28]. Targeting moieties such as small molecules, proteins, peptides, and antibodies are commonly utilised for their pharmaceutical applications. Along with them, the use of artificial single-strand oligonucleotide sequences, also known as aptamers (e.g., DNA, RNA, etc.), recently gained significant attention due to their high affinity and specificity when binding with biological molecules. 
As a diagnosis tool, iodine and lutetium radioisotopes co-emit a γ photon, which can be detected by single-photon emission computed tomography (SPECT) [29]. In addition, positron emission tomography (PET) is capable of detecting β+ decay with the emission of a positron. Further, radioisotopes can be explored as potential therapeutic agents because they have properties that enable them to emit α particles, β particles, and Auger electrons, which can damage deoxyribonucleic acid (DNA) and lead to cell apoptosis [30][31]. Among commonly used radioisotopes in biomedical applications are the short-lived β+ emitters 15O, 13N, 11C, and 18F, and the long-lived β+ emitters 89Zr, 64Cu, and 52Mn [32]. In the case of α-emitters, 213Bi, 223Ra, 211At, and 225Ac have been investigated for the treatment of cancer metastases due to their apoptotic effects [33]. Various theragnostic strategies involving them have been studied, and some radioisotopes, such as 225Ac, have been utilised for cancer therapy in preclinical and clinical trials [34][35][36][37].

2. Various Radionuclides for Theragnosis

Diagnostic biomarkers are often coupled with therapeutic medicines because they share common targets in cancer cells or tissues. Radioactive tracers or radionuclides are the pillars of this theragnostic concept, which is also employed in precision oncology. Nuclear theragnostic agents can deliver ionization and radiation to affect tumour cells or tissue, allowing for the imaging of the targets with the use of radioactive substances. In general, radiation with α and β particles is used in nuclear medicine, as both can cause severe damage to the cells due to their ability to destroy DNA. Another class of radiotherapy uses Auger electrons, which are electrons with very low energy emitted by radionuclides. The energy that is placed in nanometre–micrometre reserves induces high linear energy transfer (LET). For this reason, when it is discharged in a close propinquity to cancer cells, it can cause immense impairment by attacking DNA, both unswervingly and meanderingly through water radiolysis [38]. These radioisotopes, which are used in medicines, are studied thoroughly before being incorporated into clinical trials.
Holmium-166 (166Ho)-labelled microspheres with high activity are utilised in interventional radioembolization, along with SPECT (single-photon emission computed tomography) imaging, for dosimetry purposes. The ensuing high count rate may have an impact on dead time, lowering the dosimetric precision and image quality [39]. High purity 166Ho radioisotopes can be synthesised by one neutron activation of the 165Ho isotope. The 166Ho isotope emits two very high-energy β particles (1774.32 keV, yield 48.8%; 1854.9 keV, yield 49.9%) and γ rays (80.57 keV, yield 6.7%; 1379.40 keV, yield 0.9%). Due to its high energy, it can be visualised using a gamma camera when injected into the body [40]. 166Ho is one of the most promising radionuclides for theragnostic use due to its relatively high specific activity and short physical half-life.
Furthermore, 166Ho can be applied in theragnosis by combining it with various polymers. A relatively well-established study by Ha et al. focused on the 166Ho-chitosan complex and evaluated its theragnostic effect in 22 cystic brain tumour patients [41].
Lutetium-177 (177Lu) is a special nuclide that has gamma and beta co-emitting properties [42]. This particular property is what makes 177Lu a promising cancer theragnostic. The emission of gamma rays by 177Lu allows for a visualization of the drug’s distribution in the body due to its low energy [43]; this is called scintillation. As a moderate energy beta emitter, 177Lu has been reported to be more effective in treating small tumours than other nuclides (especially Yttrium-90, which is a stronger energy beta emitter) [44]. One of the important factors in applying nuclides to patients is the half-life. When the half-life is too long, the nuclide stays in the patient’s body for a long time, increasing the time required for hospitalization and treatment [45]
Samarium-153 (153Sm) is one of the nuclides that produces high radio-nuclidic purity through the neutron bombardment of isotopically enriched 152Sm2O3 [46]. 153Sm decays into a stable daughter nuclide, which is 153Eu, which has a half-life of 1.9 days [46]. 153Sm also has co-emitting properties with beta particles (Emax = 705 keV, 635 keV) and gamma photons (103 keV), which are suitable not only for therapy but also for SPECT imaging [47]. 153Sm-EDTMP (Quadramet®, Lantheus Medical Imaging, Inc., Billerica, MA, USA) has been approved by the US FDA as an excellent radionuclide for bone metastasis.
Yttrium-90 (90Y) is one of the most clinically used nuclides due to its unique properties. 90Y emits beta particles with 0.937 MeV in energy, and the emitted beta particles can penetrate approximately 2.5 mm of tissue [48]. It is a nuclide suitable for use in transarterial radioembolization (TARE) because it emits relatively high energy and does not penetrate deeply into tissues, resulting in a low risk of side effects. In addition, 90Y has a short half-life (2.675 days) [49]. Due to these properties, it has been widely studied and used as a cancer theragnostic, especially for liver cancer [50]

3. Radiopharmaceuticals with Radiolabelled Peptides towards Cancer Therapy: Mechanistic Pathway and Biological Paraphernalia

The neutralization of cancer cells by radiolabelled peptides follows the radiation-induced killing pathway in therapy. Radiopharmaceuticals are one of the most important tools for fighting cancer, and the development of peptide receptor radionuclide therapy (PRRT) depends on the knowledge gathered from radiotherapy [51]. The mechanistic pathways of radiopharmaceuticals show that some important parameters need to be addressed, such as the complete measurement of physiological and biological functions at the targeted sites [52]. PRRT solely depends on the process of amassing radiopharmaceuticals at the definite site, which is regulated by several biological processes such as the transference of biochemical species and enzymatic exchanges.

3.1. Radiation Dosimetry

In radiotherapy, dosimetry plays a crucial role as it reports the biological data regarding the dose absorbed by both the tumour and the normal tissue. Patient-specific dosimetry depends on the time-ordered distribution of radiopharmaceuticals, and imaging techniques are used for internal dosimetry evaluation. When the radionuclide sits on the surface of cancer cells, a small amount of released energy is deposited into the target cells [53]. This is somewhat countered by the higher concentration that may be achieved in tiny clusters of cells as compared to large, quantifiable tumours. Different techniques can be used to obtain data on the distribution of radiopharmaceuticals in a patient, which include single-photon emission computed tomography (SPECT) and positron emission tomography (PET) [54], as well as whole-body emission counting [55] and planar γ-imaging [56].
Monte Carlo (MC) simulation is a statistical method that determines the three-dimensional interactions of radioactive particles that involve a random pathway [57]. Considering tissue penetration depth, energy loss, bremsstrahlung photons, and cross-fire dosage, the MC model is typically detailed. The key benefits of MC simulations include their capacity to take into consideration an inhomogeneous radioactivity distribution, patient-specific organ geometries, induction of secondary particles (typically γ-radiation), and transitions between tissue types [58]. From measurement-based calculations to pencil beam algorithms to superposition/convolution algorithms, the dose calculation engines used for treatment planning in radiotherapy have continuously improved in terms of accuracy. Monte Carlo treatment planning (MCTP) was first developed in the 1990s, after the successful implementation of MC codes to derive patient-specific dose distribution data, but the main obstacle to efficacious clinical enactment has always been computer power [59]
The Society for Nuclear Medicine’s Medical Internal Radiation Dose (MIRD) committee devised a way to calculate the average radiation doses received by patients through radiopharmaceuticals. When utilising S values, as specified in MIRD brochures no. 5 and no. 11, it is presumed that radioactivity is distributed uniformly within organs, and that organ mass is standardised [60]. In the past, dosimetry analysis has relied on straightforward mathematical humanoid models that assume the presence of infinite homogeneous fluids with soft tissue density and include spheres of various volumes. The most recent voxel-based anthropomorphic phantoms from the MIRD/ICRP (International Commission on Radiological Protection) are designed for men, women, and children of various ages. Although diagnostic imaging may be used to create patient-specific organ masses, it is currently not possible to modify the location, tissue inhomogeneity, or form of the organs [61][62]

3.2. Localization Pathways

The term “passive diffusion” is used to describe the random movement of molecules from areas of greater to lower concentration in order to achieve homogeneity. However, in a living system, this kind of motion often involves molecular transport across a membrane. Molecule mobility across membranes is affected by factors such as pH, ionization, molecule size, and lipid solubility [63]. Since phospholipids, glycolipids, sphingolipids, and sterols are the prevalent types of lipids that make up membranes—of which phospholipids are the main component—lipid solubility is the main decisive factor. As a result, only molecules that are soluble in lipids (known as lipophilic molecules) can pierce through membranes; polar hydrophilic ones cannot. The membranes prevent the diffusion of lipophobic molecules while permitting the passage of lipophilic ones. This barrier can be broken as a result of some physiological disorders, which then allow hydrophilic molecules to diffuse into the brain tissues. 99mTc-DTPA usually follows this kind of localization mechanism when used for brain imaging. Due to its hydrophilic nature, it cannot usually cross the barrier, but when abnormalities arise which create a perturbation across the blood–brain barrier (BBB), using this passive diffusion process 99mTc-DTPA crosses the barrier [64][65][66][67].
Compartmentalised localisation is the process by which one or more preferred species are disseminated inside a limited area. In a radiopharmaceutical context, “compartment-localisation” refers to the process of confining a radiotracer within a defined volume and maintaining it long enough to allow for a thorough examination of the volume. Under normal conditions, the fluids in the body’s compartments flow in a systematic pattern, but pathologic changes can disrupt this flow, producing abnormalities. In human anatomy, the vascular system, cerebrospinal fluid space, peritoneal cavity, etc., are designated as biological compartments. 111In-DTPA imaging is used to monitor the leakage of CSF [68], which is attributed to the mechanistic pathway of unexpected leaking from its designated repository caused by pathologic alterations.

4. Fabrication of Amino Chains for Radiopharmaceutical Applications

There is an increasing need to develop advanced and sophisticated procedures for successful delivery to target sites. With advances in chemistry, chelation has been chosen as a potential method for the insertion of metallic cations (e.g., 177Lu +3, 90Y +3, etc.), considering the rate of metal complex formation and dissociation. The size of the cavity of bifunctional chelating agents (BFCAs) and the ionic radius of the cationic metal should be compatible in terms of metal binding with high stability and limiting dissociation [69]. Additionally, to circumvent the possible interference between the active site of the chelator and the receptor-binding site, a linker may be required [70]. Commonly used linkers such as PEG and amino chains have been utilised as pharmacokinetic modifiers. Many well-designed BFCAs have been explored for the fabrication of radiolabelled peptides. The common structures of acyclic and cyclic BFCAs for the development of radiolabelled amino chains are shown in Figure 1 [71].
Figure 1. Structural formulae of different chelators and co-ligands for the radiolabelling of peptides and proteins.

5. Radiolabelled Peptides Used in Cancer Theragnosis

5.1. Somatostatin Receptor-Targeted Anticancer Therapy

Somatostatin, a cyclic peptide, is a significant physiological modulator of neuroendocrine activity in various organ systems across the human anatomy. Its activity is mediated by five distinct SST receptor subtypes (SSTR1–5), and it inhibits the release of hormones as well as the growth of tumours. SSTR subtype expression fluctuates amongst various pituitary adenomas and tumours that secrete the same hormone [72]. Neuroendocrine tumours (NETs) are a category of malignancies that develop from neuroendocrine cells that are widely spread throughout the body and share similar characteristics with both endocrine (hormone-producing) and nerve cells (neurons) [73]. Numerous neuroendocrine cells, such as anterior pituitary somatotroph, thyroid C, and pancreatic islet cells, have been shown to express somatostatin receptors [74][75]. Among them, SST-2R is typically overexpressed in NETs, thus allowing for the development of theragnostics that selectively target SST-2R-positive NETs. This was the crucial piece of information that facilitated the first successful imaging of NETs, which was carried out by E.P. Krenning and colleagues [76]. They developed the first radiolabelled somatostatin analogue, namely, 123I-labelled Tyr3-octreotide. By 1990, it had been used on hundreds of patients in clinical trials, producing fine single-photon emission computed tomography (SPECT) images of localised carcinoid tumours, paragangliomas, and pancreatic endocrine tumours. However, the high biliary excretion of 123I resulted in a concentrated colonic accrual, which disrupted the construal of planar and SPECT images of tumours in the abdomen. This paved the way for the discovery of diethylenetriaminepentaacetic acid-d-phenylalanine (DTPA)-octreotide radiolabelled with indium-111 (111In-DTPA-D-Phe1-octreotide) [77], which showed high sensitivity towards the localisation of NETs.
Another commercially available SSTR analogue named 99mTc-depreotide (commercial name: NeoTect), which binds with SSTR subtypes 2, 3, and 5, is used to treat patients with non-Hodgkin’s lymphoma. Following the path of Octeroscan ™, this functions in the same way by identifying the lymphoma sites. Due to its in vivo half-life, optimal biodistribution, and high binding affinity, Depreotide (cyclo-[(N-Me)Phe-Tyr-D-Trp-Lys-Val-Hcy]CH2-CO.β-Dap-Lys-Cys-Lys.amide, P829) labelled with a beta emitter was developed as a tumour-imaging radiopharmaceutical by D.L. Bushnell and colleagues [78].
Among other somatostatin analogues, 177Lu-DOTATATE is one of the most widely used PRRTs globally. Surprisingly, when a dosimetric study was performed by introducing an albumin-binding moiety, Evans’ Blue, 177Lu-DOTATATE showed remarkably higher uptake and retention in NETs [79]. Based on the pharmacokinetic profile, it was evident that 2 h after the injection of 177Lu-DOTA-EB-TATE (177Lu-1, 4, 7, 10-tetra-azacyclododecane-1, 4, 7, 10-tetraacetic acid-Evans blue-octreotate), there was high accumulation in the blood as well as a moderate uptake in the liver, spleen, and kidneys; however, in the case of 177Lu-DOTATATE, there was surprisingly no blood accumulation detected [80]
In addition, an important study published by Thomas L. Andersen and colleagues compared [55Co]Co-DOTATATE, [64Cu]Cu-DOTATATE and [68Ga]Ga-DOTATATE to in vivo imaging characteristics in an SSTR-positive xenograft mouse model. The capacity of PET imaging to improve picture contrast and, hence, the detectability of cancers is dependent not only on the discovery of new targeting mechanisms with new tracers, but also on the imaging properties of the radionuclide. 

5.2. CD13-Targeted Anticancer Therapy

Neo-angiogenesis in the tumour stroma recruits new blood vessels from the prior vasculature. This is a multi-step process that involves growth factors, adhesion molecules, and cellular receptors, and it is crucial for tumour cell survival, proliferation, and invasion [81]. CD13, a zinc-dependent membrane-bound ectopeptidase, plays a critical role in angiogenesis [82]; it is usually overexpressed in lung, breast, and prostate cancer.
Another research work was published in 2020 by Adrienn and colleagues in which the 68Ga-NODAGA-c(NGR) [68Ga-c[Lys(1,4,7-Triazacyclononane,1-glutaric acid-4,7-acetic acid)-Asn-Gly-Arg-Glu]-NH2] peptide was used to evaluate the efficacy of in vivo molecular imaging. The aim of that research was to determine the anti-tumour effects of bestatin and actinonin treatment in subcutaneously transplanted HT1080 and B16-F10 with the use of tumour-bearing animal models. Five days after injecting bestatin and actinonin, the PET scans showed that bestatin- and actinonin-treated B16-F10 tumours exhibited significantly low radiopharmaceutical uptake and accumulation. This means that the CD13 inhibitor may be suitable for suppressing the neoangiogenic process [83]
Due to its high recurrence and mortality, ovarian cancer is a serious threat to women’s health. It is one of the most common gynaecological malignancies. Most patients with ovarian cancer show advanced intraperitoneal metastasis of the disease at prognosis, owing to its clinically silent nature [84], and the differential overexpression of CD13 has been well-documented [85]. With regard to the expression of APN/CD13, Yi Yang and colleagues reported on the use of the 68Ga-labelled dimeric cNGR peptide DOTA-c(NGR)2 [DOTA, 1,4,7,10-tetraazacyclododecane-N,N′,N″,N‴-tetraacetic acid] for the micro-PET imaging of ovarian cancer xenografts. A higher uptake of 68Ga was found in ES2 cells when compared with SKOV3 cells, which makes their compound a candidate for the assessment of CD13 expression in ovarian cancer [86]. In addition, many therapeutics have been delivered to active tumour neovascular tissue by peptides bearing the NGR motif as a homing agent. However, the affinity, specificity, and pharmacokinetics of radiolabelled NGR peptides from different molecular scaffolds and chelators vary greatly. Peptide dimerization targeting NGRs showed promise for the enhancement of pharmacokinetic properties and the tumour-to-nontarget ratio.

6. Radiolabelled Peptides Translated into Clinical Trials

Metastatic, castration-resistant prostate cancer is one of the most challenging curable diseases. Since castration-resistant prostate cancer does not respond to hormone therapy, there is an urgent need to develop a therapeutic agent that targets this type of cancer [87]. Recently, specific proteins, such as SWI/SNF, which are overexpressed in castration-resistant prostate cancer have been discovered [88]. Prostate-specific membrane antigen (PSMA) is also one of the proteins that is highly expressed in metastatic, castration-resistant prostate cancer [89]. In the clinical trial for the curing of this disease, 177Lu-PSMA-617 was investigated as a radioligand therapy that could deliver beta-particle radiation to PSMA-expressing cells and the surrounding microenvironment (NCT03511664) [89]. Oliver et al. evaluated the theragnostic efficacy of 177Lu-PSMA-617 through an international, open-label, phase 3 trial involving patients previously treated for metastatic, castration-resistant prostate cancer. The patients had undergone treatment of not only one androgen receptor pathway inhibitor, but also of one or two taxane regimens. Furthermore, PET-CT-scanned patients who used PSMA-positive gallium-68 (68Ga)-labelled PSMA-11 also participated. In the clinical trials, one group of patients additionally received 177Lu-PSMA-617 treatment (7.4 GBq every six weeks for four to six cycles) as they were provided with protocol-permitted standard care that excluded radium-223 (223Ra). The other group of patients only received standard care. The endpoints were determined based on bioimaging of the patients or the objective responses. The side effects that occurred during the clinical trial were limited to those that occurred within 30 days of the last administration of 177Lu-PSMA-617 and those that occurred before the subsequent anticancer treatment. According to the imaging-based results for progression-free survival and overall survival, 177Lu-PSMA-617 plus standard care led to significant improvement as compared with standard care alone (in the case of imaging-based progression-free survival, median was 8.7 vs. 3.4 months; hazard ratio for progression or death, 0.40; 99.2% confidence interval, 0.29 to 0.57; p < 0.001; in the case of survival, median was 15.3 vs. 11.3 months; hazard ratio for death, 0.62; 95% CI, 0.52 to 0.74; p < 0.001). However, the incidence of adverse events of grade 3 and above was higher with the administration of 177Lu-PSMA-617 (52.7%) than without (38.0%). Fortunately, the adverse events that occurred did not cause negative effects such as a reduction in the quality of life of the patients. Therefore, standard care plus radioligand therapy with 177Lu -PSMA-617 as a combination therapy significantly prolonged the imaging-based progression-free survival and overall survival of patients with PSMA-positive, metastatic, castration-resistant prostate cancer.

References

  1. Simón, M.; Jørgensen, J.T.; Khare, H.A.; Christensen, C.; Nielsen, C.H.; Kjaer, A. Combination of Lu-DOTA-TATE Targeted Radionuclide Therapy and Photothermal Therapy as a Promising Approach for Cancer Treatment: In Vivo Studies in a Human Xenograft Mouse Model. Pharmaceutics 2022, 14, 1284.
  2. Frantellizzi, V.; Verrina, V.; Raso, C.; Pontico, M.; Petronella, F.; Bertana, V.; Ballesio, A.; Marasso, S.L.; Miglietta, S.; Rosa, P.; et al. 99mTc-Labeled Keratin Gold-Nanoparticles in a Nephron-like Microfluidic Chip for Photo-Thermal Therapy Applications. Mater. Today Adv. 2022, 16, 100286.
  3. Cheng, M.H.Y.; Overchuk, M.; Rajora, M.A.; Lou, J.W.H.; Chen, Y.; Pomper, M.G.; Chen, J.; Zheng, G. Targeted Theranostic 111 In/Lu-Nanotexaphyrin for SPECT Imaging and Photodynamic Therapy. Mol. Pharm. 2022, 19, 1803–1813.
  4. Kumar, D.S.; Girija, A.R. Bionanotechnology in Cancer; Jenny Stanford Publishing: New York, NY, USA, 2022; ISBN 9780429422911.
  5. Ferreira, C.A.; Heidari, P.; Ataeinia, B.; Sinevici, N.; Sise, M.E.; Colvin, R.B.; Wehrenberg-Klee, E.; Mahmood, U. Non-Invasive Detection of Immunotherapy-Induced Adverse Events. Clin. Cancer Res. 2021, 27, 5353–5364.
  6. Luo, Q.; Zhang, Y.; Wang, Z.; Sun, Y.; Shi, L.; Yu, Y.; Shi, J.; Hu, Z.; Wang, F. A Novel Peptide-Based Probe 99mTc-PEG6-RD-PDP2 for the Molecular Imaging of Tumor PD-L2 Expression. Chin. Chem. Lett. 2022, 33, 3497–3501.
  7. Wen, X.; Zeng, X.; Liu, J.; Zhang, Y.; Shi, C.; Wu, X.; Zhuang, R.; Chen, X.; Zhang, X.; Guo, Z. Synergism of 64 Cu-Labeled RGD with Anti-PD-L1 Immunotherapy for the Long-Acting Antitumor Effect. Bioconjug. Chem. 2022, 33, 2170–2179.
  8. He, Z.; Jia, H.; Zheng, M.; Wang, H.; Yang, W.; Gao, L.; Zhang, Z.; Xue, J.; Xu, B.; Yang, W.; et al. Trp2 Peptide-Assembled Nanoparticles with Intrinsically Self-Chelating 64 Cu Properties for PET Imaging Tracking and Dendritic Cell-Based Immunotherapy against Melanoma. ACS Appl. Bio Mater. 2021, 4, 5707–5716.
  9. Wang, C.; Tian, Y.; Wu, B.; Cheng, W. Recent Progress Toward Imaging Application of Multifunction Sonosensitizers in Sonodynamic Therapy. Int. J. Nanomed. 2022, 17, 3511–3529.
  10. Dhaini, B.; Kenzhebayeva, B.; Ben-Mihoub, A.; Gries, M.; Acherar, S.; Baros, F.; Thomas, N.; Daouk, J.; Schohn, H.; Hamieh, T.; et al. Peptide-Conjugated Nanoparticles for Targeted Photodynamic Therapy. Nanophotonics 2021, 10, 3089–3134.
  11. Wang, H.; Wang, Z.; Chen, W.; Wang, W.; Shi, W.; Chen, J.; Hang, Y.; Song, J.; Xiao, X.; Dai, Z. Self-Assembly of Photosensitive and Radiotherapeutic Peptide for Combined Photodynamic-Radio Cancer Therapy with Intracellular Delivery of MiRNA-139-5p. Bioorg. Med. Chem. 2021, 44, 116305.
  12. Ho, J.A.; Wang, L.-S.; Chuang, M.-C. Nanotheranostics—A Review of Recent Publications. Int. J. Nanomed. 2012, 7, 4679–4695.
  13. Pan, T.; Mawlawi, O. PET/CT in Radiation Oncology. Med. Phys. 2008, 35, 4955–4966.
  14. Guerra Liberal, F.D.C.; Tavares, A.A.S.; Tavares, J.M.R.S. Palliative Treatment of Metastatic Bone Pain with Radiopharmaceuticals: A Perspective beyond Strontium-89 and Samarium-153. Appl. Radiat. Isot. 2016, 110, 87–99.
  15. Manzzini Calegaro, J.U.; de Podestá Haje, D.; Machado, J.; Sayago, M.; de Landa, D.C. Synovectomy Using Samarium-153 Hydroxyapatite in the Elbows and Ankles of Patients with Hemophilic Arthropathy. World J. Nucl. Med. 2018, 17, 6–11.
  16. Kolesnikov-Gauthier, H.; Lemoine, N.; Tresch-Bruneel, E.; Olivier, A.; Oudoux, A.; Penel, N. Efficacy and Safety of 153Sm-EDTMP as Treatment of Painful Bone Metastasis: A Large Single-Center Study. Support. Care Cancer 2018, 26, 751–758.
  17. Ma, H.; Li, F.; Shen, G.; Cai, H.; Liu, W.; Lan, T.; Yang, Y.; Yang, J.; Liao, J.; Liu, N. Synthesis and Preliminary Evaluation of 131 I-Labeled FAPI Tracers for Cancer Theranostics. Mol. Pharm. 2021, 18, 4179–4187.
  18. Giannakenas, C.; Kalofonos, H.P.; Apostolopoulos, D.J.; Zarakovitis, J.; Kosmas, C.; Vassilakos, P.J. Preliminary Results of the Use of Re-186-HEDP for Palliation of Pain in Patients with Metastatic Bone Disease. Am. J. Clin. Oncol. Cancer Clin. Trials 2000, 23, 83–88.
  19. Klett, R.; Lange, U.; Haas, H.; Voth, M.; Pinkert, J. Radiosynoviorthesis of Medium-Sized Joints with Rhenium-186-Sulphide Colloid: A Review of the Literature. Rheumatology 2007, 46, 1531–1537.
  20. Hertz, S. Radioactive Iodine in the Study of Thyroid Physiology. J. Am. Med. Assoc. 1946, 131, 81.
  21. Herrero Álvarez, N.; Bauer, D.; Hernández-Gil, J.; Lewis, J.S. Recent Advances in Radiometals for Combined Imaging and Therapy in Cancer. ChemMedChem 2021, 16, 2909–2941.
  22. Roll, W.; Weckesser, M.; Seifert, R.; Bodei, L.; Rahbar, K. Imaging and Liquid Biopsy in the Prediction and Evaluation of Response to PRRT in Neuroendocrine Tumors: Implications for Patient Management. Eur. J. Nucl. Med. Mol. Imaging 2021, 48, 4016–4027.
  23. Ambrosini, V.; Kunikowska, J.; Baudin, E.; Bodei, L.; Bouvier, C.; Capdevila, J.; Cremonesi, M.; de Herder, W.W.; Dromain, C.; Falconi, M.; et al. Consensus on Molecular Imaging and Theranostics in Neuroendocrine Neoplasms. Eur. J. Cancer 2021, 146, 56–73.
  24. Werner, R.A.; Solnes, L.B.; Javadi, M.S.; Weich, A.; Gorin, M.A.; Pienta, K.J.; Higuchi, T.; Buck, A.K.; Pomper, M.G.; Rowe, S.P.; et al. SSTR-RADS Version 1.0 as a Reporting System for SSTR PET Imaging and Selection of Potential PRRT Candidates: A Proposed Standardization Framework. J. Nucl. Med. 2018, 59, 1085–1091.
  25. Parghane, R.; Mitra, A.; Bannore, T.; Rakshit, S.; Banerjee, S.; Basu, S. Initial Clinical Evaluation of Indigenous 90Y-DOTATATE in Sequential Duo-PRRT Approach (177Lu-DOTATATE and 90Y-DOTATATE) in Neuroendocrine Tumors with Large Bulky Disease: Observation on Tolerability,90Y-DOTATATE Post- PRRT Imaging Characteristics (bremsstrahlung and PETCT) and early adverse effects. World J. Nucl. Med. 2021, 20, 73–81.
  26. Ngam, P.I.; Tan, E.; Lim, G.; Yan, S.X. Improving Yittrium-90 PET Scan Image Quality through Optimized Reconstruction Algorithms. J. Nucl. Med. Technol. 2023, 51, 26–31.
  27. Konuparamban, A.; Nautiyal, A.; Jha, A.; Srichandan, T.; Mithun, S.; RANGARAJAN, V. Feasibility and Reliability Assessment of Single Imaging Time-Point for Organ and Tumour Dosimetry Following 177Lu-DOTATATE PRRT. J. Nucl. Med. 2022, 63, 2817.
  28. Veenstra, E.B.; Brouwers, A.H.; de Groot, D.J.A.; Hofland, J.; Walenkamp, A.M.E.; Brabander, T.; Zandee, W.T.; Noordzij, W. Comparison of DOPA and DOTA-TOC as a PET Imaging Tracer before Peptide Receptor Radionuclide Therapy. Eur. J. Hybrid Imaging 2022, 6, 12.
  29. Wright, C.L.; Zhang, J.; Tweedle, M.F.; Knopp, M.V.; Hall, N.C. Theranostic Imaging of Yttrium-90. BioMed Res. Int. 2015, 2015, 481279.
  30. Yong, K.; Milenic, D.; Baidoo, K.; Brechbiel, M. Mechanisms of Cell Killing Response from Low Linear Energy Transfer (LET) Radiation Originating from 177Lu Radioimmunotherapy Targeting Disseminated Intraperitoneal Tumor Xenografts. Int. J. Mol. Sci. 2016, 17, 736.
  31. Kassis, A.I. Therapeutic Radionuclides: Biophysical and Radiobiologic Principles. Semin. Nucl. Med. 2008, 38, 358–366.
  32. Graves, S.A.; Hernandez, R.; Fonslet, J.; England, C.G.; Valdovinos, H.F.; Ellison, P.A.; Barnhart, T.E.; Elema, D.R.; Theuer, C.P.; Cai, W.; et al. Novel Preparation Methods of 52 Mn for ImmunoPET Imaging. Bioconjug. Chem. 2015, 26, 2118–2124.
  33. McDevitt, M.R.; Sgouros, G.; Finn, R.D.; Humm, J.L.; Jurcic, J.G.; Larson, S.M.; Scheinberg, D.A. Radioimmunotherapy with Alpha-Emitting Nuclides. Eur. J. Nucl. Med. Mol. Imaging 1998, 25, 1341–1351.
  34. McDevitt, M.R.; Ma, D.; Lai, L.T.; Simon, J.; Borchardt, P.; Frank, R.K.; Wu, K.; Pellegrini, V.; Curcio, M.J.; Miederer, M.; et al. Tumor Therapy with Targeted Atomic Nanogenerators. Science 2001, 294, 1537–1540.
  35. Müller, C.; van der Meulen, N.P.; Benešová, M.; Schibli, R. Therapeutic Radiometals Beyond 177 Lu and 90 Y: Production and Application of Promising α-Particle, β-Particle, and Auger Electron Emitters. J. Nucl. Med. 2017, 58, 91S–96S.
  36. Ling, S.W.; de Blois, E.; Hooijman, E.; van der Veldt, A.; Brabander, T. Advances in 177Lu-PSMA and 225Ac-PSMA Radionuclide Therapy for Metastatic Castration-Resistant Prostate Cancer. Pharmaceutics 2022, 14, 2166.
  37. Filippi, L.; Chiaravalloti, A.; Schillaci, O.; Bagni, O. The Potential of PSMA-Targeted Alpha Therapy in the Management of Prostate Cancer. Expert Rev. Anticancer Ther. 2020, 20, 823–829.
  38. Ku, A.; Facca, V.J.; Cai, Z.; Reilly, R.M. Auger Electrons for Cancer Therapy—A Review. EJNMMI Radiopharm. Chem. 2019, 4, 27.
  39. Stella, M.; Braat, A.J.A.T.; Lam, M.G.E.H.; de Jong, H.W.A.M.; van Rooij, R. Gamma Camera Characterization at High Holmium-166 Activity in Liver Radioembolization. EJNMMI Phys. 2021, 8, 22.
  40. Klaassen, N.J.M.; Arntz, M.J.; Gil Arranja, A.; Roosen, J.; Nijsen, J.F.W. The Various Therapeutic Applications of the Medical Isotope Holmium-166: A Narrative Review. EJNMMI Radiopharm. Chem. 2019, 4, 19.
  41. Ha, E.J.; Gwak, H.-S.; Rhee, C.H.; Youn, S.M.; Choi, C.-W.; Cheon, G.J. Intracavitary Radiation Therapy for Recurrent Cystic Brain Tumors with Holmium-166-Chico: A Pilot Study. J. Korean Neurosurg. Soc. 2013, 54, 175.
  42. Bhusari, P.; Vatsa, R.; Singh, G.; Parmar, M.; Bal, A.; Dhawan, D.K.; Mittal, B.R.; Shukla, J. Development of Lu-177-Trastuzumab for Radioimmunotherapy of HER2 Expressing Breast Cancer and Its Feasibility Assessment in Breast Cancer Patients. Int. J. Cancer 2017, 140, 938–947.
  43. Emmett, L.; Willowson, K.; Violet, J.; Shin, J.; Blanksby, A.; Lee, J. Lutetium 177 PSMA Radionuclide Therapy for Men with Prostate Cancer: A Review of the Current Literature and Discussion of Practical Aspects of Therapy. J. Med. Radiat. Sci. 2017, 64, 52–60.
  44. Dash, A.; Knapp, F.F.; Pillai, M. Targeted Radionuclide Therapy—An Overview. Curr. Radiopharm. 2013, 6, 152–180.
  45. Yeong, C.-H.; Cheng, M.; Ng, K.-H. Therapeutic Radionuclides in Nuclear Medicine: Current and Future Prospects. J. Zhejiang Univ. Sci. B 2014, 15, 845–863.
  46. Sartor, O. Overview of Samarium Sm 153 Lexidronam in the Treatment of Painful Metastatic Bone Disease. Rev. Urol. 2004, 6 (Suppl. 10), S3–S12.
  47. Eary, J.F.; Collins, C.; Stabin, M.; Vernon, C.; Petersdorf, S.; Baker, M.; Hartnett, S.; Ferency, S.; Addison, S.J.; Appelbaum, F. Samarium-153-EDTMP Biodistribution and Dosimetry Estimation. J. Nucl. Med. 1993, 34, 1031–1036.
  48. D’Arienzo, M. Emission of Β+ Particles Via Internal Pair Production in the 0+–0+ Transition of 90Zr: Historical Background and Current Applications in Nuclear Medicine Imaging. Atoms 2013, 1, 2–12.
  49. Kim, Y.-C.; Kim, Y.-H.; Uhm, S.-H.; Seo, Y.S.; Park, E.-K.; Oh, S.-Y.; Jeong, E.; Lee, S.; Choe, J.-G. Radiation Safety Issues in Y-90 Microsphere Selective Hepatic Radioembolization Therapy: Possible Radiation Exposure from the Patients. Nucl. Med. Mol. Imaging 2010, 44, 252–260.
  50. Tong, A.K.T.; Kao, Y.H.; Too, C.W.; Chin, K.F.W.; Ng, D.C.E.; Chow, P.K.H. Yttrium-90 Hepatic Radioembolization: Clinical Review and Current Techniques in Interventional Radiology and Personalized Dosimetry. Br. J. Radiol. 2016, 89, 20150943.
  51. Kong, G.; Hicks, R.J. PRRT for Higher-Grade Neuroendocrine Neoplasms: What Is Still Acceptable? Curr. Opin. Pharmacol. 2022, 67, 102293.
  52. Desai, P.; Rimal, R.; Sahnoun, S.E.M.; Mottaghy, F.M.; Möller, M.; Morgenroth, A.; Singh, S. Radiolabeled Nanocarriers as Theranostics—Advancement from Peptides to Nanocarriers. Small 2022, 18, 2200673.
  53. Gudkov, S.; Shilyagina, N.; Vodeneev, V.; Zvyagin, A. Targeted Radionuclide Therapy of Human Tumors. Int. J. Mol. Sci. 2015, 17, 33.
  54. Lee, S.J.; Park, H.J. Single Photon Emission Computed Tomography (SPECT) or Positron Emission Tomography (PET) Imaging for Radiotherapy Planning in Patients with Lung Cancer: A Meta-Analysis. Sci. Rep. 2020, 10, 14864.
  55. Park, M.; Kim, M.; Yoo, J.; Jo, M. Validation of the Whole-Body Counting Measurement in a Radiation Emergency. Appl. Radiat. Isot. 2021, 168, 109476.
  56. Tashima, H.; Yamaya, T. Compton Imaging for Medical Applications. Radiol. Phys. Technol. 2022, 15, 187–205.
  57. Andreo, P. Monte Carlo Simulations in Radiotherapy Dosimetry. Radiat. Oncol. 2018, 13, 121.
  58. Lee, C.; Park, B.; Lee, S.-S.; Kim, J.-E.; Han, S.-S.; Huh, K.-H.; Yi, W.-J.; Heo, M.-S.; Choi, S.-C. Efficacy of the Monte Carlo Method and Dose Reduction Strategies in Paediatric Panoramic Radiography. Sci. Rep. 2019, 9, 9691.
  59. Muraro, S.; Battistoni, G.; Kraan, A.C. Challenges in Monte Carlo Simulations as Clinical and Research Tool in Particle Therapy: A Review. Front. Phys. 2020, 8, 567800.
  60. Snyder, W.S.; Fisher, H.L.; Ford, M.R.; Warner, G.G. Estimates of Absorbed Fractions for Monoenergetic Photon Sources Uniformly Distributed in Various Organs of a Heterogeneous Phantom. J. Nucl. Med. 1969, 10 (Suppl. S3), 7–52.
  61. Dieudonné, A.; Hobbs, R.F.; Bolch, W.E.; Sgouros, G.; Gardin, I. Fine-Resolution Voxel S Values for Constructing Absorbed Dose Distributions at Variable Voxel Size. J. Nucl. Med. 2010, 51, 1600–1607.
  62. Siegel, J.A.; Thomas, S.R.; Stubbs, J.B.; Stabin, M.G.; Hays, M.T.; Koral, K.F.; Robertson, J.S.; Howell, R.W.; Wessels, B.W.; Fisher, D.R.; et al. MIRD Pamphlet No. 16: Techniques for Quantitative Radiopharmaceutical Biodistribution Data Acquisition and Analysis for Use in Human Radiation Dose Estimates. J. Nucl. Med. 1999, 40, 37S–61S.
  63. Amaya, H. Diffusion Processes in Tumors: A Nuclear Medicine Approach. In AIP Conference Proceedings; AIP Publishing LLC: Melville, NY, USA, 2016; p. 080015.
  64. Rani, N.; Singh, B.; Kumar, N.; Singh, P.; Hazari, P.P.; Vyas, S.; Hooda, M.; Chitkara, A.; Shekhawat, A.S.; Gupta, S.K.; et al. -Bis-Methionine-DTPA Single-Photon Emission Computed Tomography Impacting Glioma Management: A Sensitive Indicator for Postsurgical/Chemoradiotherapy Response Assessment. Cancer Biother. Radiopharm. 2021, 36, 568–578.
  65. Verma, A.; Hesterman, J.Y.; Chazen, J.L.; Holt, R.; Connolly, P.; Horky, L.; Vallabhajosula, S.; Mozley, P.D. Intrathecal 99m Tc-DTPA Imaging of Molecular Passage from Lumbar Cerebrospinal Fluid to Brain and Periphery in Humans. Alzheimer’s Dement. Diagn. Assess. Dis. Monit. 2020, 12, e12030.
  66. Santra, A.; Sharma, P.; Kumar, R. Use of 99m-Technetium-Glucoheptonate as a Tracer for Brain Tumor Imaging: An Overview of Its Strengths and Pitfalls. Indian J. Nucl. Med. 2015, 30, 1–8.
  67. Kung, H.F. New Technetium 99m-Labeled Brain Perfusion Imaging Agents. Semin. Nucl. Med. 1990, 20, 150–158.
  68. Porubcin, S.; Rovnakova, A.; Zahornacky, O.; Jarcuska, P. Diagnostic Value of Radioisotope Cisternography Using 111In-DTPA in a Patient with Rhinorrhea and Purulent Meningitis. Medicina 2022, 58, 714.
  69. Wangler, B.; Schirrmacher, R.; Bartenstein, P.; Wangler, C. Chelating Agents and Their Use in Radiopharmaceutical Sciences. Mini-Rev. Med. Chem. 2011, 11, 968–983.
  70. Okarvi, S.M. Recent Developments in 99Tcm-Labelled Peptide-Based Radiopharmaceuticals. Nucl. Med. Commun. 1999, 20, 1093–1112.
  71. Dijkgraaf, I.; Agten, S.M.; Bauwens, M.; Hackeng, T.M. Strategies for Site-Specific Radiolabeling of Peptides and Proteins. In Radiopharmaceuticals—Current Research for Better Diagnosis and Therapy; IntechOpen: London, UK, 2022.
  72. Bronstein, M.D. Acromegaly: Molecular Expression of Somatostatin Receptor Subtypes and Treatment Outcome. In Pituitary Today: Molecular, Physiological and Clinical Aspects; KARGER: Basel, Switzerland, 2006; pp. 129–134.
  73. de Herder, W.W.; Rehfeld, J.F.; Kidd, M.; Modlin, I.M. A Short History of Neuroendocrine Tumours and Their Peptide Hormones. Best Pract. Res. Clin. Endocrinol. Metab. 2016, 30, 3–17.
  74. Bauer, W.; Briner, U.; Doepfner, W.; Haller, R.; Huguenin, R.; Marbach, P.; Petcher, T.J.; Pless, J. SMS 201–995: A Very Potent and Selective Octapeptide Analogue of Somatostatin with Prolonged Action. Life Sci. 1982, 31, 1133–1140.
  75. Reubi, J.C.; Maurer, R. Autoradiographic Mapping of Somatostatin Receptors in the Rat Central Nervous System and Pituitary. Neuroscience 1985, 15, 1183–1193.
  76. Lamberts, S.W.; Reubi, J.C.; Bakker, W.H.; Krenning, E.P. Somatostatin Receptor Imaging with 123I-Tyr3-Octreotide. Z. Gastroenterol. 1990, 28 (Suppl. S2), 20–21.
  77. Bakker, W.H.; Albert, R.; Bruns, C.; Breeman, W.A.P.; Hofland, L.J.; Marbach, P.; Pless, J.; Pralet, D.; Stolz, B.; Koper, J.W.; et al. -Octreotide, a Potential Radiopharmaceutical for Imaging of Somatostatin Receptor-Positive Tumors: Synthesis, Radiolabeling and in Vitro Validation. Life Sci. 1991, 49, 1583–1591.
  78. Bushnell, D.L.; Menda, Y.; Madsen, M.T.; Link, B.K.; Kahn, D.; Truhlar, S.M.; Juweid, M.; Shannon, M.; Murguia, J.S. 99mTc-Depreotide Tumour Uptake in Patients with Non-Hodgkin’s Lymphoma. Nucl. Med. Commun. 2004, 25, 839–843.
  79. Liu, Q.; Cheng, Y.; Zang, J.; Sui, H.; Wang, H.; Jacobson, O.; Zhu, Z.; Chen, X. Dose Escalation of an Evans Blue–Modified Radiolabeled Somatostatin Analog 177Lu-DOTA-EB-TATE in the Treatment of Metastatic Neuroendocrine Tumors. Eur. J. Nucl. Med. Mol. Imaging 2020, 47, 947–957.
  80. Zhang, J.; Wang, H.; Jacobson, O.; Cheng, Y.; Niu, G.; Li, F.; Bai, C.; Zhu, Z.; Chen, X. Safety, Pharmacokinetics, and Dosimetry of a Long-Acting Radiolabeled Somatostatin Analog 177 Lu-DOTA-EB-TATE in Patients with Advanced Metastatic Neuroendocrine Tumors. J. Nucl. Med. 2018, 59, 1699–1705.
  81. Alessandrini, L.; Ferrari, M.; Taboni, S.; Sbaraglia, M.; Franz, L.; Saccardo, T.; Del Forno, B.M.; Agugiaro, F.; Frigo, A.C.; Dei Tos, A.P.; et al. Tumor-Stroma Ratio, Neoangiogenesis and Prognosis in Laryngeal Carcinoma. A Pilot Study on Preoperative Biopsies and Matched Surgical Specimens. Oral Oncol. 2022, 132, 105982.
  82. Mina-Osorio, P. The Moonlighting Enzyme CD13: Old and New Functions to Target. Trends Mol. Med. 2008, 14, 361–371.
  83. Kis, A.; Dénes, N.; Szabó, J.P.; Arató, V.; Beke, L.; Matolay, O.; Enyedi, K.N.; Méhes, G.; Mező, G.; Bai, P.; et al. In Vivo Molecular Imaging of the Efficacy of Aminopeptidase N (APN/CD13) Receptor Inhibitor Treatment on Experimental Tumors Using 68Ga-NODAGA-c(NGR) Peptide. Biomed Res. Int. 2021, 2021, 6642973.
  84. Surowiak, P.; Drąg, M.; Materna, V.; Suchocki, S.; Grzywa, R.; Spaczyński, M.; Dietel, M.; Oleksyszyn, J.; Zabel, M.; Lage, H. Expression of Aminopeptidase N/CD13 in Human Ovarian Cancers. Int. J. Gynecol. Cancer 2006, 16, 1783–1788.
  85. Meng, Y.; Zhang, Z.; Liu, K.; Ye, L.; Liang, Y.; Gu, W. Aminopeptidase N (CD13) Targeted MR and NIRF Dual-Modal Imaging of Ovarian Tumor Xenograft. Mater. Sci. Eng. C 2018, 93, 968–974.
  86. Yang, Y.; Zhang, J.; Zou, H.; Shen, Y.; Deng, S.; Wu, Y. Synthesis and Evaluation of 68 Ga-Labeled Dimeric CNGR Peptide for PET Imaging of CD13 Expression with Ovarian Cancer Xenograft. J. Cancer 2021, 12, 244–252.
  87. Crowley, F.; Sterpi, M.; Buckley, C.; Margetich, L.; Handa, S.; Dovey, Z. A Review of the Pathophysiological Mechanisms Underlying Castration-Resistant Prostate Cancer. Res. Rep. Urol. 2021, 13, 457–472.
  88. Cyrta, J.; Augspach, A.; De Filippo, M.R.; Prandi, D.; Thienger, P.; Benelli, M.; Cooley, V.; Bareja, R.; Wilkes, D.; Chae, S.-S.; et al. Role of Specialized Composition of SWI/SNF Complexes in Prostate Cancer Lineage Plasticity. Nat. Commun. 2020, 11, 5549.
  89. Sartor, O.; de Bono, J.; Chi, K.N.; Fizazi, K.; Herrmann, K.; Rahbar, K.; Tagawa, S.T.; Nordquist, L.T.; Vaishampayan, N.; El-Haddad, G.; et al. Lutetium-177–PSMA-617 for Metastatic Castration-Resistant Prostate Cancer. N. Engl. J. Med. 2021, 385, 1091–1103.
More
Information
Subjects: Oncology
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , ,
View Times: 246
Revisions: 2 times (View History)
Update Date: 03 May 2023
1000/1000
ScholarVision Creations