IMR Press / FBL / Volume 24 / Issue 4 / DOI: 10.2741/4742
Review
GCPII and its close homolog GCPIII: from a neuropeptidase to a cancer marker and beyond
Show Less
1 Institute of Organic Chemistry and Biochemistry of the Czech Academy of Sciences, Flemingovo n. 2, Prague 6, 16610, Czech Republic
2 First Faculty of Medicine, Charles University, Katerinska 32, Prague 2, 12108, Czech Republic
3 Department of Biochemistry, Faculty of Science, Charles University, Albertov 6, Prague 2, 12843, Czech Republic
*Correspondence: jan.konvalinka@uochb.cas.cz (Jan Konvalinka)
Front. Biosci. (Landmark Ed) 2019, 24(4), 648–687; https://doi.org/10.2741/4742
Published: 1 March 2019
(This article belongs to the Special Issue Dipeptidyl peptidase IV and related molecules in health and disease)
Abstract

Glutamate carboxypeptidases II and III (GCPII and GCPIII) are highly homologous di-zinc metallopeptidases belonging to the M28 family. These enzymes are expressed in a variety of tissues, including the brain, prostate, kidney, testis and jejunum. GCPII has been recognized as a neuropeptidase in the central nervous system, as a folate hydrolase participating in absorption of folates in the jejunum and, most importantly, as a prostate-specific membrane antigen that is highly expressed in prostate adenocarcinoma. Furthermore, it has been identified in the neovasculature of most human solid tumors. In contrast, GCPIII has not been associated with any specific physiological function or pathology, and its expression, activity and inhibition have not been as well-studied. In this review, we provide an overview of the current understanding of the structure, enzymatic activity, substrate specificity, and tissue distribution of these two homologous enzymes. We discuss their potential physiological functions and describe the available animal models, including genetically modified mice. We also review the potential use of specific monoclonal antibodies and small-molecule inhibitors recognizing GCPII/III for diagnosis, imaging and experimental therapy of human cancers and other pathologies.

Keywords
GCPII
GCPIII
PSMA
FOLH1
NAALAD2
NAAG
Folate
BCG
Review
2. INTRODUCTION

Metallopeptidases require bivalent metallic cations for catalysis. According to the MEROPS database (1), metallopeptidases can be divided into 99 peptidase families based on statistically significant similarities in their amino acid sequences. Each family is typically characterized by additional features. For example, the M28 peptidase family contains aminopeptidases and carboxypeptidases that require co-catalytic zinc for their enzymatic activity. This family consists of six subfamilies, with subfamily M28B containing glutamate carboxypeptidase II (GCPII) and glutamate carboxypeptidase III (GCPIII).

GCPII and GCPIII are type II transmembrane glycoproteins with a short intracellular portion, a single membrane-spanning segment, and a large extracellular domain with carboxypeptidase activity (Figure 1) (2-4). In the prostate, GCPII is referred to as prostate-specific membrane antigen (PSMA), and is highly expressed in prostate adenocarcinoma (5). It also has been identified in the neovasculature of most solid tumors (6). GCPII has been independently described in the brain as N-acetylated alpha-linked acidic dipeptidase (NAALADase) and in jejunum as a folate hydrolase responsible for absorption of folates (3, 7-9). While the physiological function of GCPII has been well-characterized in the nervous system and small intestine, it is still not fully understood in other highly GCPII-expressing tissues such as the prostate and kidney. In contrast to the wealth of information about GCPII, only a few reports describing GCPIII have been published to date. Its physiological function has yet to be fully elucidated, and there is no information available on its possible involvement in human diseases.

Figure 1

Comparison of the overall 3D structures of GCPII and GCPIII. A: Structural alignment of GCPII (PDB 2PVW) and GCPIII (PDB 3FF3) monomers. Both monomers are depicted in cartoon representation. The GCPIII monomer is colored cyan, and the GCPII monomer is colored wheat. C- and N-termini of GCPII monomer are indicated. B: Crystal structure of GCPII homodimer in cartoon representation. One monomeric unit is shown in wheat while the other is colored according to the domain organization: the protease-like domain in red, the apical domain in yellow, and the C-terminal domain in orange. The zinc cations are shown as blue spheres, the chloride anion as a green sphere, and the calcium cation as a violet sphere. The carbohydrate moieties are shown as sticks in cyan. C: Comparison of the surface electrostatic potentials of the GCPII (PDB 2PVW) and GCPIII (PDB 3FF3) X-ray structures. Two representations of both structures are shown rotated by 180° (the left orientation being identical to the orientation of the molecule in panel A). The electrostatic gradient is shown from the most negatively charged surface (red) to the most positively charged surface (blue). The figures were created using The PyMOL Molecular Graphics System, Version 0.99 Schrödinger, LLC. The electrostatic potential maps were calculated using PyMol APBS Tools.

The genes encoding GCPII and GCPIII map to chromosome 11, and their exon/intron structures are similar (10). While the FOLH1 gene encoding GCPII is localized at position 11p11-p12 (2, 11, 12), the NAALAD2 gene encoding GCPIII is at position 11q14.3.-q21 (13). GCPII and GCPIII share 67% amino acid sequence identity and 81% sequence similarity. The overall 3D structures of these proteins are very similar, with a root-mean-square deviation (RMSD) for 676 aligned C-alpha atoms of 0.59 Å (see Figure 1A) (13, 14). The extracellular parts of both proteins form homodimers in which each monomer folds into three domains—the protease-like domain, the apical domain and the C-terminal or dimerization domain (see Figure 1B) (14-17). Both proteins are post-translationally N-glycosylated, which is crucial for their enzymatic activities (18-21).

Despite these many similarities, several structural differences between GCPII and GCPIII have been identified. These include differences in surface charge distribution that could be a major determinant in physiologically relevant interactions with hypothetical protein partners or ligands (see Figure 1C) and slightly divergent active sites that result in distinct substrate specificities and catalytic efficiencies (14).

This review aims to compare the enzymatic activities and expression profiles of GCPII and GCPIII under physiological and pathophysiological conditions. We describe known functions of GCPII in the brain and small intestine and discuss the potential roles of GCPII and GCPIII in these and other tissues. Finally, we provide an overview of the use of GCPII as a diagnostic and therapeutic target, describe the current state of inhibitor development, and discuss why it is of high interest to identify small-molecule compounds that selectively bind GCPII but not GCPIII.

3. GCPII AND GCPIII ENZYMATIC ACTIVITIES

As their names suggest, both GCPII and GCPIII are hydrolases capable of cleaving a C-terminal glutamate from their substrates. Indeed, both enzymes can process most N-acetylated dipeptides of the form Ac-X-Glu-OH, where X is any proteinogenic amino acid (21-23). The catalytic efficiencies of GCPII and GCPIII differ significantly for some dipeptides (21), which is not surprising because there are several important structural differences in the active sites of these enzymes.

First, while the occupancy of zinc ions within the GCPII active site reaches 100%, zinc ions within the GCPIII active site display much lower occupancies (80-95% for Zn1 and 40-80% for Zn2) (see Figure 2A) (14, 16). This is likely caused by substitution of Asn519 in GCPII with Ser509 in GCPIII. Ser509 has two different conformations in the GCPIII structure. In the first conformation, Ser509 interacts with the Zn2-coordinating amino acid Asp443 in a similar matter as Asn519 interacts with Asp543 in GCPII. However, in the second conformation, Ser509 does not interact with Asp443, enabling Asp443 to adopt an alternative conformation in which it does not coordinate the Zn2 ion. This conformational flexibility is likely the reason for the low occupancy of Zn2 within the GCPIII active site (see Figure 2A) (14, 24).

Figure 2

Structural differences between the GCPII and GCPIII active sites. A: Comparison of the zinc coordination sites of GCPII (PDB 2OR4) and GCPIII (PDB 3FF3). For GCPIII, two different conformational states (A and B) are shown. The amino acids that adopt different conformations in the GCPIII structure are colored yellow, while the rest of the amino acids are colored gray. All amino acids are shown in stick representation. Oxygen atoms are colored red, and nitrogen atoms blue. Zinc ions are shown as blue spheres and the chloride ion as a green sphere. Water molecules are shown as red spheres. The hydrogen bonds and zinc co-ordinations are shown as black dashed lines. B: Highlight of the arene-binding site (ABS) within GCPII (PDB 2XEG). GCPII is shown in surface representation colored gray with a portion of the molecule omitted for visual clarity. The ligand ARM-P4 is shown as sticks and colored yellow. Trp541 and Arg511, the amino acids comprising the ABS, are shown as sticks and colored cyan and magenta, respectively. Oxygen atoms are colored red, and nitrogen atoms blue. Zinc ions are shown as green spheres. The figures were created using The PyMOL Molecular Graphics System, Version 0.9.9 Schrödinger, LLC.

Second, GCPII contains a so-called arene-binding site (ABS) that is absent in GCPIII (24-26). This exosite is formed by the Trp541 and Arg511 side chains at the entrance of a tunnel leading into the GCPII active site (see Figure 2B) and seems to be important for both processing of some endogenous substrates and binding of highly specific inhibitors (25, 26). Indeed, pi-pi stacking interactions with the Trp541 and Arg511 side chains participate in the binding of arene-containing ligands into the GCPII active site. In GCPIII, Trp541 is substituted by Lys531, which effectively prevents binding of similar ligands to the enzyme, as the Lys side chain is unable to engage in pi-pi stacking interactions.

To date, two physiologically relevant substrates processed by both GCPII and GCPIII have been identified—N-acetyl-L-aspartyl-L-glutamate (NAAG) and polyglutamylated folates (FolGlun) (13, 24). In addition, both enzymes have been shown to cleave naturally occurring β-citryl-L-glutamate (BCG), although the catalytic efficiency of GCPII is extremely low, thus making BCG a specific substrate of GCPIII (24).

3.1. NAAG hydrolyzing activity

Both GCPII and GCPIII cleave NAAG, yielding N-acetyl-L-aspartate and L-glutamate (Figure 3). NAAG hydrolyzing activity was first detected in 1984 on membranes isolated from rat brain cortex (27). Soon after, this activity was assigned to a membrane-bound metallopeptidase termed N-acetylated alpha-linked acidic dipeptidase (NAALADase) (28, 29), which was later designated as GCPII (8, 9, 30). In 1999, a second membrane-bound enzyme capable of NAAG hydrolysis was identified (13). This enzyme was first named NAALADase II but later designated as GCPIII (4). Since then, several studies comparing the catalytic efficiencies of GCPII and GCPIII have been published (4, 20, 21, 24).

Figure 3

NAAG-hydrolyzing reaction catalyzed by GCPII or GCPIII.

Bzdega et al. reported that rat GCPII and mouse GCPIII have similar catalytic efficiencies and pH dependencies for NAAG hydrolysis (4). In contrast, a comparative analysis of human GCPII and GCPIII performed by Hlouchova et al. using purified recombinant proteins revealed that GCPII processes NAAG with 3- to greater than 10-fold higher catalytic efficiency than GCPIII, depending on the reaction conditions. Moreover, the pH and salt concentration optima of human GCPII and GCPIII differ significantly (21).

Interestingly, GCPII and GCPIII respond differently to the presence of divalent metal ions during the NAAG-hydrolyzing reaction. While GCPII is essentially metal-insensitive, GCPIII NAAG-hydrolyzing activity can be stimulated by addition of Mn2+ or Zn2+ and inhibited by addition of Ca2+ (20, 24). This difference relates to replacement of Asn519 in GCPII with Ser509 in GCPIII. Recombinant GCPII becomes metal-sensitive when a N519S mutation is introduced (20). Because the occupancy of the Zn2 ion within the GCPIII active site is low, Zn2 in the GCPIII structure can be replaced with a different metal ion, creating a heterometallic cluster and altering the enzyme’s catalytic efficiency. Notably, a reciprocal mutation in GCPIII (i.e., S509N) does not abolish the metal dependency of the enzyme for NAAG-hydrolyzing activity, suggesting that the mechanism of metal ion involvement in the reaction is likely more complex (20).

GCPII was recently shown to process the endogenous tripeptide N-acetyl-L-aspartyl-L-glutamyl-L-glutamate (NAAG2) (31). Using recombinant enzymes, we demonstrated that not only GCPII but also GCPIII is capable of NAAG2 cleavage (Navratil et al., unpublished observation). However, the enzyme kinetics for this reaction has not yet been evaluated in detail.

3.2. FolGlun hydrolyzing activity

The hydrolysis of polyglutamylated folates (FolGlun) by both GCPII and GCPIII yields folate and free L-glutamate (Figure 4) (24, 26). The role of GCPII as an intestinal folate hydrolase was identified by two independent research groups in the 1990s (7, 32), although detection of FolGlun hydrolyzing activity in intestinal mucosa dates back to 1969 (33). Our group recently reported that GCPIII also is able to cleave FolGlun, and we directly compared GCPII and GCPIII catalytic efficiencies in FolGlun-hydrolyzing reactions (24).

Figure 4

FolGlun-hydrolyzing reaction catalyzed by GCPII or GCPIII.

Interestingly, while cleavage of monoglutamylated folate (FolGlu1) by the two enzymes is comparable, the processing of polyglutamylated folates (FolGlu2-6) by GCPII is almost two orders of magnitude more efficient than by GCPIII (24, 26). This observation can be explained by the engagement of the GCPII ABS in recognizing polyglutamylated folates. The ABS binds the hydrophobic pteroate moiety, leading to potential stabilization of FolGlu2-6 in the active site. As the ABS is absent in GCPIII, FolGlu2-6 cannot be stabilized in the GCPIII active site in a similar manner. When processing FolGlu1, GCPII is unable to fully utilize the ABS for binding of the pteroate moiety of the substrate, leading to cleavage kinetics comparable to that of GCPIII.

The metal dependence of GCPII and GCPIII catalytic efficiency has only been explored for FolGlu1 cleavage, and researchers found a similar trend as for NAAG cleavage. GCPII appears insensitive to metals during FolGlu1 processing. On the other hand, GCPIII shows the highest catalytic efficiency in the presence of Zn2+, followed by Mn2+ and Ca2+ (24).

3.3. BCG hydrolyzing activity

Both GCPII and GCPIII cleave BCG, yielding citrate and L-glutamate (Figure 5) (24). BCG-hydrolyzing activity was first detected in the rat testis in 1983 (34), followed by purification of a BCG-hydrolyzing enzyme from the rat testis particulate in 1995 (35). However, identification of this BCG-hydrolase as GCPIII was published more than 15 years later (20). In their pivotal work, Collard and co-workers discovered the BCG-hydrolyzing activity of mouse GCPIII and reported that mouse GCPII is not capable of BCG cleavage. Using highly purified recombinant enzymes, we recently found that human GCPII can process BCG, but with catalytic efficiency three to five orders of magnitude lower than that of human GCPIII (24). These observations suggest that BCG can indeed be considered a specific substrate of GCPIII.

Figure 5

BCG-hydrolyzing reaction catalyzed by GCPII or GCPIII.

The reasons for such a striking difference in the ability of these enzymes to process BCG are not fully understood, but they seem to be related to replacement of Asn519 in GCPII with Ser509 in GCPIII. Indeed, Collard and co-workers showed that introducing a N519S mutation into mouse GCPII renders the enzyme capable of BCG hydrolysis. Introducing the reciprocal S509N mutation in mouse GCPIII abolishes its BCG-hydrolyzing activity (20). The main reason that substitution of Asn519 in GCPII with Ser509 in GCPIII affects BCG hydrolysis is likely connected with the fact that this substitution influences the metal ion occupancy in the active site.

Similar to the NAAG-hydrolyzing and FolGlun-hydrolyzing reactions, BCG hydrolysis by GCPIII is stimulated by addition of certain divalent metal ions, while GCPII is metal-insensitive. Interestingly, the metal-dependency of GCPIII for BCG hydrolysis shows an almost completely opposite trend than in the case of NAAG cleavage—it is stimulated by addition of Mn2+ or Ca2+ and inhibited by addition of Zn2+ (20, 24). Collard and co-workers proposed that the presence of Mn2+ or Ca2+ in the reaction leads to generation of heterometallic Zn-Mn or Zn-Ca clusters within the GCPIII active site. Because BCG brings more negative charges into the active site than NAAG, metal ions with the ability to bind more ligands (6 or 7 ligands for Ca2+ compared with 4 or 5 for Zn2+) are preferable (20). However, QM/MM calculations by Navratil et al. did not confirm that Ca2+ would be thermodynamically more favorable for BCG as a substrate; this suggests that a more complex mechanism of action underlies the metal-dependency of GCPIII enzymatic activity (24).

Mentioned discrepancies could be addressed by solving the GCPIII structure in complex with BCG and appropriate metal ions. The comparison of such a structure with the available structure of the GCPII-BCG complex, together with QM/MM/MD calculations, could address not only why BCG is a specific substrate of GCPIII but also how metal ions are involved in the reaction.

4. GCPII AND GCPIII EXPRESSION PROFILES

Since the development of the first anti-GCPII antibody, 7E11 (5), several research groups have thoroughly studied GCPII localization in human and animal tissues under physiological and pathophysiological conditions. However, the data are rather inconsistent, likely for two main reasons. First, different studies used distinct detection methods, ranging from Western blotting, ELISA, and immunohistochemistry to in vivo imaging (17). The data gained from these experiments thus may differ dramatically due to the variable detection limits of each method. Second, various antibodies have been used, and most of them were poorly characterized, which may have led to their improper use and thus generation of false positive/negative results. To provide guidance on how to use anti-GCPII antibodies in the most efficient way for individual immunochemical methods, we recently published a detailed characterization of the most commonly used anti-GCPII antibodies (36).

In contrast to the wealth of data on GCPII, almost no data are available on the distribution of GCPIII within the human body. Notably, several antibodies recognizing GCPII also detect GCPIII (36). Therefore, it is possible that some tissues reported to be GCPII-expressing are in fact also or solely GCPIII-expressing. The major challenge facing determination of GCPIII localization has been the absence of an available antibody selective for native or denatured GCPIII. However, this obstacle was recently overcome by use of a BCG-cleavage assay for specific GCPIII detection in human tissues (24).

4.1. Expression of GCPII and GCPIII in healthy tissues

Despite inconsistencies in the data, a general consensus has been reached on GCPII expression within the human body: it is expressed in the prostate (5, 37-48), nervous system (39, 45-47, 49), small intestine (39-41, 43-45, 47, 48), and kidney (5, 37, 40, 41, 43, 45-47, 50). Closer examinations of individual tissues using either human or rat histological samples localized GCPII to the acinar epithelium in the prostate (40, 45, 47), astrocytes and Schwann cells in the nervous system (49, 51-53), duodenal mucosa in the small intestine (40, 45), and proximal tubules in the kidney (40, 45, 47, 50). In addition, GCPII has been repeatedly detected in human blood (54-57), and we have observed pronounced GCPII expression in the salivary glands (unpublished observation). Several studies also suggest that GCPII expression is not restricted to the sites mentioned above, but that the enzyme is to a lesser extent ubiquitous throughout the human body, expressed in tissues such as the breast, colon, heart, liver, ovary, testis, urinary bladder, and uterus. (44-47).

Only one report has addressed localization of GCPIII in humans (24). Among the tissues examined (testis, ovary, placenta, spleen, prostate, pancreas, heart, jejunum, kidney, and brain), GCPIII was detected mainly in the testis and to a much lesser extent in the prostate, brain, kidney, and small intestine. Nevertheless, a more rigorous investigation using a broader set of human tissues and more biological replicates is needed.

4.2. Expression of GCPII in pathology

While GCPIII has not been associated with any pathophysiological condition to date, GCPII has been linked to several different diseases. In particular, GCPII is recognized as a prostate cancer marker because its high expression within prostate tumor cells has been reported repeatedly (5, 37-41, 43-46, 58-61). Moreover, the level of GCPII expression increases with increasing prostate cancer grade (58-60), and a substantial amount of GCPII is present in different prostate cancer metastases (5, 38, 40, 59, 62).

Although the first studies assessing GCPII expression in malignancies suggested that it is expressed only in prostate cancer (5, 37, 39), it has since been detected in numerous other cancers. These include bladder carcinoma, breast carcinoma, colorectal adenocarcinoma, non-small cell lung carcinoma, glioma, clear cell renal cell carcinoma, and thyroid carcinoma (for a complete list, see Table 1) (40, 41, 43-46, 50, 63-85). Nevertheless, while GCPII is localized mainly in prostate cancer tumor cells (38, 40, 58-60), its expression in most other cancers is restricted to the tumor-associated neovasculature (40, 41, 43, 50, 66, 86).

Table 1 Expression of GCPII in cancers other than prostate cancer

Recently, overexpression of GCPII was reported in inflammatory bowel disease (IBD) (87). FOLH1 gene upregulation in IBD had been demonstrated previously (88-90), and GCPII expression was confirmed in one case using immunohistochemistry (89). In the recent study, researchers explored whether GCPII is present in an active form in the affected intestinal mucosa of IBD patients. They reported a 3- to 40-fold increase in GCPII enzymatic activity compared to colon samples from healthy volunteers (87). The significance of this increase remains to be resolved but the findings suggest that GCPII could serve as a biomarker for IBD.

5. PHYSIOLOGICAL FUNCTIONS OF GCPII AND GCPIII

Although GCPII is expressed in many tissues, its function has been established only in the brain and small intestine. The functions in both tissues are distinct but closely correlate with the known enzymatic activities of GCPII. To study the physiological role of GCPII in more detail, several independent research groups attempted to generate GCPII null mutant mice, with conflicting results. While some reported that GCPII deficiency is embryonically lethal, others demonstrated that inactivation of the Folh1 gene generates viable GCPII null mutant mice with no obvious phenotype (91-96). Investigations of viable GCPII null mutant mice have focused mainly on the nervous system (91, 94, 97, 98). Further examination of other tissues is thus desirable, as it could confirm or disprove the non-enzyme functions of GCPII suggested in the literature, as well as point out yet-to-be-revealed physiological roles of GCPII.

The physiological function of GCPIII has not yet been determined. Because GCPIII possesses similar enzyme activities as GCPII, it has been suggested that GCPIII serves as an enzyme complementary to GCPII. This hypothesis was further strengthened by the fact that NAAG-hydrolyzing activity was detected in the brains of GCPII null mutant mice (91). Nevertheless, considering that the expression level of GCPIII in the testis is more than one order of magnitude higher than that of GCPII, it is highly probable that GCPIII also fulfills other functions in the human body (24).

5.1. Established function of GCPII in the brain

In the brain, GCPII functions as a neuromodular enzyme by cleaving and thus deactivating the most abundant peptide neurotransmitter, NAAG (99). During synaptic transmission, NAAG is released together with other neurotransmitters from the presynaptic neuron into the synaptic cleft and participates in downregulation of postsynaptic neuron stimulation (99-101). This regulation may be executed through two different types of receptors capable of NAAG binding: N-methyl-D-aspartate receptors (NMDARs) and metabotropic glutamate receptors (mGluRs) (102, 103).

The physiological relevance of the interaction of NAAG with NMDARs has been debated in the literature (104). Moreover, a clear consensus has not been reached on which action within the synaptic cleft this interaction would evoke (105). On the other hand, the pathway of NAAG action through mGluRs has been established and seems to be generally accepted (104, 105) (Figure 6). Once released into the synaptic cleft, NAAG activates type 3 metabotropic glutamate receptors (mGluR3) located at the membranes of both presynaptic neurons and astrocytes (106-108). The subsequent events depend on the site of action. While stimulation of presynaptic mGluR3 inhibits additional release of neurotransmitters from the presynaptic neuron (109-112), activation of mGluR3 on astrocytes triggers a signaling pathway that results in the production and secretion of neuroprotective growth factors (113-115). Both these actions help prevent excessive nerve stimulation that could lead to pathological effects. Because GCPII serves as a NAAG deactivator, it is not surprising that substantial efforts have been made to identify a selective GCPII inhibitor for potential use in the treatment of neurological disorders (for more information see Section 6.2.).

Figure 6

Proposed functions of NAAG, glutamate and GCPII on synaptic terminals. The neurotransmitters NAAG and glutamate are released into the synaptic cleft from a presynaptic neuron. Glutamate activates N-methyl-D-aspartate receptors (NMDAR), which leads to apoptosis under pathological conditions associated with excessive glutamate levels. NAAG activates metabotropic glutamate 3 receptors (mGluR3) on presynaptic neurons (resulting in the decrease of released glutamate) and on astrocytes (leading to the secretion of neuroprotective factors, such as TGF-β). GCPII (and possibly GCPIII) hydrolyzes NAAG into N-acetyl-L-aspartyl (NAA) and glutamate and consequently abolishes the protective effects of NAAG.

5.2. Established function of GCPII in the small intestine

In the small intestine, GCPII is involved in folate absorption. Natural folates occur in the diet as poly-gamma-glutamates (116-118). The process of natural food folate absorption comprises two basic steps. First, GCPII located in the brush border membrane of the small intestine consecutively cleaves polyglutamylated folates to produce monoglutamylated forms of folate (119). Second, monoglutamylated folates are taken up by intestinal cells through the proton-coupled folate transporter (120). This transporter has been shown to be highly specific for monoglutamylated folates (121). GCPII thus plays a crucial role in folate absorption. For this reason, several research groups have studied the possible influence of FOLH1 gene polymorphisms on folate metabolism, but the data are rather conflicting (122-135).

One of the two naturally occurring polymorphisms reported to date—rs61886492, also known as C1561T—has been associated with increased serum folate levels (123, 126, 127, 130-132). While two reports also showed an association between rs61886492 and decreased serum homocysteine levels (130, 132), others failed to confirm this link (123, 126, 127, 131). Moreover, several studies did not observe any effect of rs61886492 on serum folate or homocysteine (124, 125, 128, 129). At the protein level, the rs61886492 mutation results in the H475Y mutation (122), which does not alter GCPII activity towards polyglutamylated folates (26). If this polymorphism is indeed involved in altered serum folate levels, the mechanism of action does not directly relate to the enzymatic activity and remains to be determined.

The second naturally occurring FOLH1 polymorphism—rs202676, also known as T484C—has been studied to a much lesser extent. Nevertheless, recent reports have shown that this polymorphism correlates with lower red blood cell folate levels (133-135), which is in accordance with the predicted hypofunction of GCPII caused by this missense mutation (133). Indeed, structural modeling suggested that the corresponding protein mutation Y75H leads to alteration in the secondary and tertiary structure of GCPII and thus affects the binding of polyglutamylated folates (133).

Numerous studies have addressed the possible association of FOLH1 gene polymorphisms with the incidence of various pathological conditions resulting from folate metabolism dysregulation. These include severity of negative symptoms in schizophrenia (136), neural tube defects (125, 126, 133, 137-140), risk for different cancers (128, 132, 141-144), and cardiovascular disease risk (123). No association with colorectal (128, 132, 141), gastric (142), and lung (143) cancer risk or cardiovascular disease risk (123) was observed. Interestingly, one study suggested that rs61886492 might have a protective role in breast cancer; however, the cohort consisted of only 68 women from one ethnic group (144). Neural tube defects are the best-studied pathological condition in relation to FOLH1 gene polymorphism. However, studies of this correlation are rather inconsistent since one report showed that FOLH1 gene polymorphism is potential risk factor for neural tube defects (133) while others failed to find this association (125, 126, 137, 139, 140). Further research is thus necessary to shed the light on the possible involvement of FOLH1 gene polymorphisms in neural tube defects.

5.3. Null mutant mice

The first GCPII null mutant mice were described in 2002 by Bacich and co-workers (91). These mice were prepared by deletion of intron-exon boundary sequences of exons 1 and 2 of the Folh1 gene and insertion of several in-frame stop codons between exon 1 and 2. The resulting GCPII null mutant mice expressed neither a GCPII mRNA nor a shortened version of the protein, and NAAG cleavage activity was dramatically decreased in the brain and kidney of these mice compared to wild-type (WT) mice. Interestingly, the levels of NAAG and glutamate in the brain were not significantly altered in the GCPII null mutant mice. The authors thus hypothesized that deletion of the Folh1 gene may induce changes in expression of other genes, including Naalad2 (which encodes GCPIII). Phenotypic examination of the GCPII null mutant mice did not reveal any significant differences in standard neurological behavior in comparison with WT mice. Nevertheless, these GCPII null mutant mice were protected from peripheral neuropathies and traumatic brain injury (TBI) (97).

In 2003, scientists from the research group of Joseph T. Coyle reported that manipulation of embryonic stem cells by deletion of exons 9 and 10 of the Folh1 gene leads to embryonic lethality (92). The same group subsequently tried to reproduce the work of Bacich and co-workers by deleting exons 1 and 2 of the Folh1 gene, but their attempt did not lead to generation of viable GCPII null mutant mice (93).

In 2015, another research group reported preparation of GCPII null mutant mice by deleting exons 3 to 5 of the Folh1 gene (94). The resulting GCPII null mutant mice showed similar phenotypic characteristics as the mice prepared by Bacich and co-workers. While no overt behavioral differences between GCPII null mutant and WT mice were observed, GCPII null mutant mice were less susceptible to TBI and showed improved long-term behavioral outcomes after TBI (94).

Basic information about the generation and phenotyping of GCPII null mutant mice also can be found on the International Mouse Phenotyping Consortium (IMPC) website (95). These viable GCPII null mutant mice were prepared by deleting exon 3 of the Folh1 gene. In contrast to the results of others, GCPII null mutant mice characterized by the IMPC do not show any clear phenotype (95). However, the full set of phenotyping data has not yet been released.

Recently, we also reported generation of viable GCPII null mutant mice using TALEN-mediated disruption of the sequence encoding GCPII active site (96). It remains unclear why neither strategy attempted by Coyle’s research group led to production of live GCPII null mutant mice. However, thanks to the work of four other laboratories, it now seems to be evident that knock-out of the GCPII-encoding gene does not lead to mouse embryonic lethality.

No GCPIII null mutant mice have been reported to date, although generation and investigation of such mice could be very useful. It could not only shed the light on the physiological role of GCPIII but also point out unique physiological roles of GCPII.

5.4. Suggested functions of GCPII and GCPIII

Outside of the nervous system and small intestine, GCPII’s role in human physiology remains elusive. Researchers are debating whether GCPII may have other, unrecognized physiological substrates or serve as a receptor for a yet-to-be-identified extracellular ligand. However, not enough data has been published for any hypothesis to be widely accepted by the scientific community. The receptor theory is based on the fact that GCPII possesses sequence and structural similarity to the transferrin receptor (16, 145) and is capable of internalization upon ligand binding (146, 147). One report even suggested that GCPII itself might be able to transport folate into cells (148). However, no follow-up research has been performed to date. The description of GCPII as a folate transporter, which has recently repeatedly appeared in the literature (6, 140, 149), should thus be avoided.

Although GCPII is present in prostate cancer cells and in the neovasculature of many solid tumors, its function in these pathologies is still unclear. Some reports link GCPII to tumor progression and carcinogenesis (148, 150), in accordance with its expression in prostate cancer metastases (5, 38, 40, 59, 62). On the other hand, others suggest that GCPII instead suppresses cancer invasiveness (151). A detailed recent review of GCPII’s potential involvement in malignant processes is available elsewhere (6).

While many studies have aimed to elucidate GCPII function, the physiological role of GCPIII has not been extensively investigated. As GCPIII has been identified as a BCG-hydrolyzing enzyme, it is likely involved in BCG metabolism. However, the physiological role of BCG has been studied by only one research group, and it remains poorly understood. Potentially, BCG might be important for brain development, as it was found in high concentrations in the newborn rat brain but disappeared with maturation (152, 153). Similarly, BCG was detected in the germinal cells of the rat testes, and its amount increased with the appearance of late spermatocytes and early spermatids (154), suggesting it could be involved in spermatogenesis. Follow-up studies linked BCG to physiologically important chelation of iron and copper (155, 156). Indeed, BCG, as a low molecular weight chelator, may play a role in reactive oxygen species-scavenging activities, inhibition of xanthine oxidase, or activation of aconitase (155-157).

6. DIAGNOSTICS AND THERAPY

Given its overexpression in pathology and its physiological functions, GCPII is considered a promising target for diagnosis and therapy of prostate and other cancers, as well as a target for treatment of neurological disorders and potentially IBD. For potential treatment of neurological disorders and IBD, researchers aim to block GCPII hydrolase activity using selective inhibitors, while in experimental cancer treatments, GCPII serves as a molecular address for targeted delivery of therapeutics. The therapeutic and diagnostic potential of GCPII was recently comprehensively reviewed by Evans and coauthors (6) and Wustemann and coauthors (158). We briefly summarize the current state of the field.

6.1. GCPII/GCPIII inhibitors

Since inhibition of NAAG-hydrolyzing activity was shown to be neuroprotective in mice, considerable effort has been made to develop potent GCPII/GCPIII inhibitors (159-164). Both enzymes contain two active-site zinc ions, and their S1’ pockets exhibit a strong preference for glutamate/glutamate-like moieties (22). This active site “architecture” led to the development of a canonical GCPII inhibitor structure consisting of a glutamate moiety (fitting into the S1’ subsite) and a zinc-binding group (“chelating” the zinc ions). Based on the nature of the zinc-binding group, the inhibitors are usually categorized into three groups: phosphorus-containing compounds (e.g. phosphonates and phosphinates), ureas, and thiol-based compounds (165, 166). Unfortunately, the majority of GCPII/GCPIII inhibitors identified to date are rather polar compounds with poor bioavailability (17, 167). There is thus a need to identify novel inhibitor scaffolds that are less polar, yet still potent and selective. Recently developed high-throughput assays could be applied for this purpose (168, 169).

Due to the similarity between the GCPII and GCPIII active sites, most inhibitors developed to date bind to both enzymes. While this feature may be beneficial for potential drugs targeting neurological disorders, it could be a pitfall for small molecule-based targeting in cancers, which needs to be highly specific. Furthermore, as the physiological role of GCPIII is unknown, inhibition of this close GCPII homolog could cause unexpected side effects. In particular, the testis, as a high GCPIII-expressing organ, may represent a possible off-target tissue for GCPII-based tumor-targeted drug delivery.

Considering the high similarity of the GCPII and GCPIII active sites, it is not surprising that development of GCPII selective inhibitors is challenging. Nevertheless, Tykvart et al. recently prepared a GCPII inhibitor with a rigid linker that exhibits more than 4,000-fold selectivity for GCPII (Ki = 49 pM) over GCPIII (Ki = 210 nM) (170). The authors also investigated some common GCPII inhibitors, such as DKFZ-PSMA-11 and DCIBzL, and found out that these inhibitors also possess a three-order-of-magnitude selectivity for GCPII over GCPIII (170). However, their Ki values towards GCPIII are in the low nanomolar range (Ki = 13 nM and 3 nM, respectively), due to which these inhibitors may be still capable of interfering with potential GCPIII functions, including BCG metabolism. Overall, the work of Tykvart et al. indicated that selectivity for GCPII over GCPIII can be achieved by developing compounds that bind to GCPII exosites that are not present in GCPIII. In addition, other structural features that differ between GCPII and GCPIII, such as the “entrance lid”, Pro-rich region, and zinc cation occupancies (14), should be considered for the future development of highly selective GCPII inhibitors.

6.2. GCPII in neurological disorders

A number of in vivo studies have shown that inhibition of GCPII ameliorates conditions caused by pathologically increased glutamate levels and have demonstrated the potential of selective GCPII inhibitors for treatment of various neurological disorders (reviewed in (165, 171)). These include ischemic and traumatic brain injury (172-175), amyotrophic lateral sclerosis (176), inflammatory and neuropathic pain (164, 177, 178), schizophrenia (179, 180), multiple sclerosis (181), and autoimmune encephalomyelitis (182). The majority of these experiments were carried out in rats and mice, which serve as suitable animal models due to the similarities between rat and mouse GCPII and their human orthologue in the context of neurological diseases (47, 183).

Importantly, GCPII inhibitors do not seem to influence normal glutamate functions. Rather, their effect is confined to suppressing excessive pathological glutamate signaling (172). This offers a potential advantage over other strategies modulating glutamatergic neurotransmission, such as use of NMDAR antagonists, which often cause undesirable side effects (184).

Although several GCPII inhibitors have been successful in animal models (e.g. 2-PMPA, ZJ-43 and 2-MPPA), none has become a clinically used therapeutic agent, in part due to generally poor oral bioavailability and low blood-brain barrier penetration. The thiol-based inhibitor 2-MPPA was evaluated in phase I clinical trials (185) that found it safe and well-tolerated; however, its further development was stopped, likely due to concerns about the adverse properties of thiols (e.g. their nucleophilicity, propensity to oxidation, metabolic instability) (186). Recently, a prodrug strategy has emerged to increase the lipophilicity of GCPII inhibitors and thus improve oral absorption. In this approach, the polar groups necessary for inhibitor binding to the GCPII active site (e.g. phosphonate, carboxylate, and hydroxamate) are masked by hydrophobic moieties (isopropyloxycarbonyloxymethyl, thiolactones) that are cleaved off in the target tissues by hydrolases (167, 187-189).

6.3. GCPII in inflammatory bowel disease

GCPII enzyme activity has been shown to be elevated in IBD patients (87). In a follow-up study, the researchers reported that systemic administration of the GCPII inhibitor 2-PMPA markedly decreases IBD severity and ameliorates symptoms in mice (87, 190). Even though the mechanism of action has not been determined, these results suggest that GCPII inhibition might contribute to IBD therapy.

6.4. GCPII in cancer

The current major focus of GCPII research is imaging and therapy of prostate cancer and its metastases. In addition, given its expression in the neovasculature of many solid tumors, GCPII is being investigated for its potential to serve as an imaging and therapeutic target in other cancers (191). Various techniques are available for molecular imaging of GCPII within malignant tissue. The most frequently used methods are radioligand-based positron emission tomography (PET) and single photon emission computed tomography (SPECT), usually combined with computed tomography (CT) (192, 193).

The only agent approved by the U.S. Food and Drug Administration (FDA) for imaging of prostate cancer is the 111In-labeled monoclonal antibody 7E11 (194, 195), known as 111In-capromab pendetide (trade name ProstaScint™). Second-generation radiolabeled monoclonal antibodies recognizing the extracellular portion of GCPII (such as J591, which enables visualization of viable cells) have been tested in clinical trials (196, 197). The main disadvantage of the antibodies for imaging is half-life being as long as several days, which may cause harm in off-target tissues. From this reason, low-molecular-weight inhibitors have been developed since they are capable of rapid removal from circulation by renal filtration. Urea-based GCPII inhibitors radiolabeled with 18F, 68Ga, and 64Cu are considered the most promising PET tracers for GCPII imaging. These include 68Ga-PSMA-11 (or 68Ga-PSMA-11-HBED-CC; Figure 7A) and 18F-DCFPyL (Figure 7B), which demonstrated potential for identification of metastases and recurrent prostate cancer (198-205). Further research and development of improved and novel GCPII imaging probes is ongoing (206-212).

Figure 7

GCPII-targeted PET imaging agents. A: Radiometal-labeled agent 68Ga-PSMA-11 (198). B: Radiolabeled agent (18F)DCFPyL (199).

Radioimmunotherapy/radioligand therapy is the most common approach for prostate cancer therapy using GCPII as the target. The first such “therapeutic” agent targeting GCPII was 90Y-7E11 antibody; however, it exhibited hematologic toxicity in clinical trials (213, 214). This toxicity, as well as the inability of 7E11 to bind viable tumor cells, spurred the development of second-generation radiolabeled antibodies such as J591 (215). 177Lu-J591 demonstrated excellent GCPII-targeting in men with metastatic castration-resistant prostate cancer (mCRPC), resulting in a dose-dependent decline in their PSA levels (197, 216, 217). Recently, the “third-generation” of highly potent monoclonal antibodies recognizing GCPII has been introduced, and these represent prime candidates for the development of novel theranostics to target GCPII (218).

Small molecules labeled with therapeutic radionuclides, such as the recently developed 177Lu-PSMA-617, are also promising therapeutic agents for treatment of mCRPC (219). 177Lu-PSMA-617 has been studied intensively during the last two years (220-226). Two recent German retrospective multicenter studies evaluating 177Lu-PSMA-617 radioligand therapy (145 and 59 mCRPC patients) showed that the radioconjugate exhibits favorable safety and efficacy (a ≥ 50% PSA decline occurred in 45% and 53% of patients, respectively) (222, 223). Similar results, showing high response rates, low toxicity and reduction of pain in men with mCRPC, were obtained by a single-arm, single-center, phase 2 trial performed in Australia (224). In February 2017, the first U.S. multicenter phase II trial of 177Lu-PSMA-617 radioligand therapy received FDA clearance (estimated study completion date: April 11, 2019) (225). Quite recently (May 2018), a multicenter randomized phase III study has been launched (estimated study completion date: May 2021), which will compare overall survival in patients with GCPII-positive mCRPC who receive 177Lu-PSMA-617 in addition to best standard of care versus patients treated with best standard of care alone (226).

In addition to radiotherapy, antibodies and small molecules could be used for delivery of cytotoxic molecules and toxins. Deimmunized J591 was conjugated to an extremely potent microtubule-depolymerizing drug, DM-1 (227); unfortunately, the conjugate exhibited neurotoxicity and limited activity in mCRPC patients (likely due to the deconjugation of DM-1) (228). More encouragingly, a conjugate of an anti-GCPII antibody and the potent antimitotic drug monomethyl auristatin E exhibited low toxicity and a ≥ 50% PSA decline in 33% of mCRPC patients (229-231). Other conjugates, including a conjugate of J591 and ricin A-chain and a conjugate of the anti-GCPII single-chain antibody fragment D7 and the toxic domain of Pseudomonas exotoxin A (PE40) (232, 233), have not yet been tested in a clinical setting. Nevertheless, the latter significantly inhibited growth of a GCPII-positive tumor xenograft in mice (234). In addition, many small selective GCPII inhibitors conjugated with cytotoxic agents have been prepared (158, 235), with one already being investigated in a phase I clinical trial for the treatment of mCRPC (236).

Other GCPII-targeted drug delivery systems include anti-GCPII aptamers (237-243), polymer-based nanocarriers (244-249), liposomes with co-entrapped drugs (250-252), and nanoparticles targeting GCPII, often designed as theranostics (253-255). We have recently developed structurally diverse nanoparticles (polymer conjugates, nanodiamonds, and virus-like particles) decorated with GCPII inhibitors, which could be used for future drug delivery (256, 257).

Finally, there is an increasing number of reports describing GCPII-based immunotherapy of prostate cancer, although only a few studies have been performed in a clinical setting. These investigational therapies are based on dendritic cells (258, 259), GCPII-CD3 diabodies and Fab conjugates (260, 261), cytotoxic T lymphocytes (262), T-cells bearing a chimeric antigen receptor (263-266), PEI-PEG-DUPA conjugates targeting polyinosine/polycytosine acid inducing tumor regression (267), and DNA-encoded anti-GCPII monoclonal antibodies mediating antibody-dependent cellular cytotoxicity (268).

7. SUMMARY AND PERSPECTIVES

Despite decades of research, the scientific community remains intensely interested in GCPII. A Pubmed query using GCPII/PSMA as a search term returns almost 800 entries from 2017 to present, most dealing with GCPII as a molecular address for prostate cancer theranostics. This approach leverages molecular recognition of GCPII by monoclonal antibodies, nanobodies, nucleic acid aptamers, or small-molecule inhibitors for the specific delivery of nanoparticles (polymer conjugates, nanodiamonds, virus-like particles) decorated by cytotoxic drugs, radiolabels, or fluorescent probes for efficient cancer treatment and imaging. Dozens of papers describing novel antibody conjugates or nanoparticle formulations are published almost every week. Regardless, very little is known about the role of GCPII in cancer development, both in the prostate and in the neovasculature of other solid tumors. The proteolytic function of GCPII is almost certainly not crucial for its putative role in cancer development, as inhibition of GCPII alone has no anticancer activity. Potentially, GCPII could serve as a receptor of a yet-to-be-identified signaling ligand contributing to cell growth or migration. Additionally, the physiological function of GCPII outside the brain and jejunum remains unknown, as does the possible role of GCPII in development and pathogenesis of IBD. Furthermore, while GCPII is highly expressed in the kidneys, there is no hypothesis for its role in this organ.

Even less is known about the function of GCPIII. It has been long viewed as a mere isoenzyme of GCPII, not distinguishable by its activity or by immunochemical methods. Only recently it was established that this enzyme is highly expressed in the testis, where it co-localizes with its most specific cognate substrate, BCG. These findings also raise the need for further research. The role of BCG and its degradation in the testis is yet to be determined, as is the molecular mechanism by which GCPIII specifically recognizes BCG while GCPII does so much less effectively. A structural analysis of the complex of GCPIII with BCG completed by QM/MM/MD calculations could help explain the specificity of the enzyme and the role of metals in this particular enzymatic activity. It is difficult to imagine that such a high concentration of an enzyme and specific cognate substrate in the same tissue does not have an important physiological consequence. To elucidate the physiological function of the BCG-hydrolyzing activity of GCPIII, specific novel tools will be required, such as GCPIII-specific antibodies or selective inhibitors of this homolog. The distinct structural features of GCPIII discussed in this review could be considered to achieve this task.

Null mutant animals likely represent the most direct avenue to addressing these unanswered questions. Some attempts to create GCPII null mutant mice resulted in embryonic lethality, while others led to viable mice with no clear phenotype. Our own animal experiments support the latter finding – GCPII null mutant mice develop normally with minor phenotypic characteristics appearing in the later adult stage. GCPIII is most likely capable of complementing the lost GCPII activity in these experimental animals. Development of GCPIII null mutant mice and GCPII/III double null mutant mice would be key to elucidating the function of both homologs. Improving our understanding of the physiological function of these fascinating enzymes will hopefully lead to their better use as tools for future diagnostic and therapeutic approaches.

8. ACKNOWLEDGEMENTS

The authors would like to acknowledge Dr. Hillary Hoffman for language editing. This work was supported by InterBioMed project LO 1302 from the Ministry of Education of the Czech Republic.

Abbreviations: ABS: arene-binding site, BCG: β-citryl-L-glutamate, FDA: Food and Drug Administration, FolGlun: polyglutamylated folates, GCPII: glutamate carboxypeptidase II, GCPIII: glutamate carboxypeptidase III, IBD: inflammatory bowel disease, IMPC: International Mouse Phenotyping Consortium, mGluRs: metabotropic glutamate receptors, mCRPC: metastatic castration-resistant prostate cancer, NAA: N-acetyl-L-aspartyl, NAAG: N-acetyl-L-aspartyl-L-glutamate, NAAG2: N-acetyl-L-aspartyl-L-glutamyl-L-glutamate, NAALADase: N-acetylated alpha-linked acidic dipeptidase, NMDARs: N-methyl-D-aspartate receptors, PET: positron emission tomography, PSMA: prostate-specific membrane antigen, TBI: traumatic brain injury, WT: wild-type

References
[1]
ND Rawlings AJ Barrett R Finn Twenty years of the MEROPS database of proteolytic enzymes, their substrates and inhibitors Nucleic Acids Res 2016 44 D1 D343 50 DOI:10.1.093/nar/gkv1118
[2]
DS O'Keefe SL Su DJ Bacich Y Horiguchi Y Luo CT Powell D Zandvliet PJ Russell PL Molloy NJ Nowak TB Shows C Mullins RA Vonder Haar WR Fair WDW Heston Mapping, genomic organization and promoter analysis of the human prostate-specific membrane antigen gene Biochimica Et Biophysica Acta-Gene Structure and Expression 1998 1443 1-2 113 127
[3]
ND Rawlings AJ Barrett Structure of membrane glutamate carboxypeptidase Bba-Protein Struct M 1997 1339 2 247 252 DOI:10.1.016/S0167-4838(97)00008-3
[4]
T Bzdega SL Crowe ER Ramadan KH Sciarretta RT Olszewski OA Ojeifo VA Rafalski B Wroblewska JH Neale The cloning and characterization of a second brain enzyme with NAAG peptidase activity J Neurochem 2004 89 3 627 635 DOI:10.1.111/j.1471-4159.2.004.0.2361.x
[5]
JS Horoszewicz E Kawinski GP Murphy Monoclonal antibodies to a new antigenic marker in epithelial prostatic cells and serum of prostatic cancer patients Anticancer Res 1987 7 5B 927 35
[6]
JC Evans M Malhotra JF Cryan CM O'Driscoll The therapeutic and diagnostic potential of the prostate specific membrane antigen/glutamate carboxypeptidase II (PSMA/GCPII) in cancer and neurological disease Br J Pharmacol 2016 173 21 3041 3079 DOI:10.1.111/bph.13576
[7]
JT Pinto BP Suffoletto TM Berzin CH Qiao S Lin WP Tong F May B Mukherjee WD Heston Prostate-specific membrane antigen: a novel folate hydrolase in human prostatic carcinoma cells Clin Cancer Res 1996 2 9 1445 51 DOI:10.1.016/S0006-8993(98)00244-3
[8]
R Luthi-Carter AK Barczak H Speno JT Coyle Hydrolysis of the neuropeptide N-acetylaspartylglutamate (NAAG) by cloned human glutamate carboxypeptidase II Brain Res 1998 795 1-2 341 8 DOI:10.1.016/S0006-8993(98)00244-3
[9]
R Luthi-Carter AK Barczak H Speno JT Coyle Molecular characterization of human brain N-acetylated alpha-linked acidic dipeptidase (NAALADase) J Pharmacol Exp Ther 1998 286 2 1020 5 DOI:10.2.174/092986712799462676
[10]
K Hlouchova V Navratil J Tykvart P Sacha J Konvalinka GCPII Variants, Paralogs and Orthologs Curr Medicinal Chem 2012 19 9 1316 1322 DOI:10.2.174/092986712799462676
[11]
CW Rinkerschaeffer AL Hawkins SL Su RS Israeli CA Griffin JT Isaacs WDW Heston Localization and Physical Mapping of the Prostate-Specific Membrane Antigen (Psm) Gene to Human-Chromosome-11 Genomics 1995 30 1 105 108 DOI:10.1.006/geno.1995.0.019
[12]
BH Maraj JP Leek M Karayi M Ali NJ Lench AF Markham Detailed genetic mapping around a putative prostate-specific membrane antigen locus on human chromosome 11p11.2 Cytogenet Cell Genet 1998 81 1 3 9 DOI:10.1.159/000014999
[13]
MN Pangalos JM Neefs M Somers P Verhasselt M Bekkers L van der Helm E Fraiponts D Ashton RD Gordon Isolation and expression of novel human glutamate carboxypeptidases with N-acetylated alpha-linked acidic dipeptidase and dipeptidyl peptidase IV activity J Biol Chem 1999 274 13 8470 8483 DOI:10.1.074/jbc.274.1.3.8.470
[14]
K Hlouchova C Barinka J Konvalinka J Lubkowski Structural insight into the evolutionary and pharmacologic homology of glutamate carboxypeptidases II and III Febs Journal 2009 276 16 4448 4462 DOI:10.1.111/j.1742-4658.2.009.0.7152.x
[15]
MI Davis MJ Bennett LM Thomas PJ Bjorkman Crystal structure of prostate-specific membrane antigen, a tumor marker and peptidase Proceedings of the National Academy of Sciences of the United States of America 2005 102 17 5981 5986 DOI:10.1.073/pnas.0502101102
[16]
JR Mesters C Barinka WX Li T Tsukamoto P Majer BS Slusher J Konvalinka R Hilgenfeld Structure of glutamate carboxypeptidase II, a drug target in neuronal damage and prostate cancer Embo J 2006 25 6 1375 1384 DOI:10.1.038/sj.emboj.7600969
[17]
C Barinka C Rojas B Slusher M Pomper Glutamate Carboxypeptidase II in Diagnosis and Treatment of Neurologic Disorders and Prostate Cancer Curr Medicinal Chem 2012 19 6 856 870 DOI:10.2.174/092986712799034888
[18]
A Ghosh WD Heston Effect of carbohydrate moieties on the folate hydrolysis activity of the prostate specific membrane antigen Prostate 2003 57 2 140 51 DOI:10.1.002/pros.10289
[19]
C Barinka P Sacha J Sklenar P Man K Bezouska BS Slusher J Konvalinka Identification of the N-glycosylation sites on glutamate carboxypeptidase II necessary for proteolytic activity Protein Sci 2004 13 6 1627 35 DOI:10.1.110/ps.04622104
[20]
F Collard D Vertommen S Constantinescu L Buts E Van Schaftingen Molecular Identification of beta-Citrylglutamate Hydrolase as Glutamate Carboxypeptidase 3 J Biol Chem 2011 286 44 38220 38230 DOI:10.1.074/jbc.M111.2.87318
[21]
K Hlouchova C Barinka V Klusak P Sacha P Mlcochova P Majer L Rulisek J Konvalinka Biochemical characterization of human glutamate carboxypeptidase III J Neurochem 2007 101 3 682 696 DOI:10.1.111/j.1471-4159.2.006.0.4341.x
[22]
C Barinka M Rinnova P Sacha C Rojas P Majer BS Slusher J Konvalinka Substrate specificity, inhibition and enzymological analysis of recombinant human glutamate carboxypeptidase II J Neurochem 2002 80 3 477 87 DOI:10.1.046/j.0022-3042.2.001.0.0715.x
[23]
J Tykvart C Barinka M Svoboda V Navratil R Soucek M Hubalek M Hradilek P Sacha J Lubkowski J Konvalinka Structural and Biochemical Characterization of a Novel Aminopeptidase from Human Intestine J Biol Chem 2015 DOI:10.1.074/jbc.M114.6.28149
[24]
M Navratil J Tykvart J Schimer P Pachl V Navratil TA Rokob K Hlouchova L Rulisek J Konvalinka Comparison of human glutamate carboxypeptidases II and III reveals their divergent substrate specificities FEBS J 2016 283 13 2528 45 DOI:10.1.111/febs.13761
[25]
AX Zhang RP Murelli C Barinka J Michel A Cocleaza WL Jorgensen J Lubkowski DA Spiegel A Remote Arene-Binding Site on Prostate Specific Membrane Antigen Revealed by Antibody-Recruiting Small Molecules Journal of the American Chemical Society 2010 132 36 12711 12716 DOI:10.1.021/Ja104591m
[26]
M Navratil J Ptacek P Sacha J Starkova J Lubkowski C Barinka J Konvalinka Structural and biochemical characterization of the folyl-poly-gamma-l-glutamate hydrolyzing activity of human glutamate carboxypeptidase II FEBS J 2014 DOI:10.1.111/febs.12857
[27]
N Riveros F Orrego A study of possible excitatory effects of N-acetylaspartylglutamate in different in vivo and in vitro brain preparations Brain Res 1984 299 2 393 5 DOI:10.1.016/0006-8993(84)90727-3
[28]
MB Robinson RD Blakely JT Coyle Quisqualate selectively inhibits a brain peptidase which cleaves N-acetyl-L-aspartyl-L-glutamate in vitro Eur J Pharmacol 1986 130 3 345 7 DOI:10.1.016/0014-2999(86)90291-8
[29]
MB Robinson RD Blakely R Couto JT Coyle Hydrolysis of the brain dipeptide N-acetyl-L-aspartyl-L-glutamate. Identification and characterization of a novel N-acetylated alpha-linked acidic dipeptidase activity from rat brain J Biol Chem 1987 262 30 14498 506
[30]
R Luthi-Carter UV Berger AK Barczak M Enna JT Coyle Isolation and expression of a rat brain cDNA encoding glutamate carboxypeptidase II Proc Natl Acad Sci U S A 1998 95 6 3215 20 DOI:10.1.073/pnas.95.6.3.2.15
[31]
J Lodder-Gadaczek I Becker V Gieselmann L Wang-Eckhardt M Eckhardt N-Acetylaspartylglutamate Synthetase II Synthesizes N-Acetylaspartylglutamylglutamate J Biol Chem 2011 286 19 16693 16706 DOI:10.1.074/jbc.M111.2.30136
[32]
CH Halsted EH Ling R Luthi-Carter JA Villanueva JM Gardner JT Coyle Folylpoly-gamma-glutamate carboxypeptidase from pig jejunum - Molecular characterization and relation to glutamate carboxypeptidase II J Biol Chem 1998 273 32 20417 20424 DOI:10.1.074/jbc.273.3.2.2.0417
[33]
IH Rosenberg RR Streiff HA Godwin WB Castle Absorption of polyglutamic folate: participation of deconjugating enzymes of the intestinal mucosa N Engl J Med 1969 280 18 985 8 DOI:10.1.056/NEJM196905012801804
[34]
M Miyake T Innami Y Kakimoto A beta-citryl-L-glutamate-hydrolysing enzyme in rat testes Biochim Biophys Acta 1983 760 2 206 14 DOI: 0304-4165(83)90165-4
[35]
M Asakura Y Nagahashi M Hamada M Kawai K Kadobayashi M Narahara S Nakagawa Y Kawai T Hama M Miyake Purification and properties of beta-citryl-L-glutamate-hydrolysing enzyme from rat testis particulate Biochim Biophys Acta 1995 1250 1 35 42 DOI:10.1.016/0167-4838(95)00077-8
[36]
J Tykvart V Navratil F Sedlak E Corey M Colombatti G Fracasso F Koukolik C Barinka P Sacha J Konvalinka Comparative analysis of monoclonal antibodies against prostate-specific membrane antigen (PSMA) Prostate 2014 74 16 1674 1690 DOI:10.1.002/Pros.22887
[37]
AD Lopes WL Davis MJ Rosenstraus AJ Uveges SC Gilman Immunohistochemical and pharmacokinetic characterization of the site-specific immunoconjugate CYT-356 derived from antiprostate monoclonal antibody 7E11-C5 Cancer Res 1990 50 19 6423 9
[38]
GL Wright Jr C Haley ML Beckett PF Schellhammer Expression of prostate-specific membrane antigen in normal, benign, and malignant prostate tissues Urol Oncol 1995 1 1 18 28
[39]
JK Troyer ML Beckett GL Wright Jr Detection and characterization of the prostate-specific membrane antigen (PSMA) in tissue extracts and body fluids Int J Cancer 1995 62 5 552 8
[40]
DA Silver I Pellicer WR Fair WDW Heston C CordonCardo Prostate-specific membrane antigen expression in normal and malignant human tissues Clinical Cancer Research 1997 3 1 81 85
[41]
H Liu P Moy S Kim Y Xia A Rajasekaran V Navarro B Knudsen NH Bander Monoclonal antibodies to the extracellular domain of prostate-specific membrane antigen also react with tumor vascular endothelium Cancer Res 1997 57 17 3629 34
[42]
S Zhang HS Zhang VE Reuter SF Slovin HI Scher PO Livingston Expression of potential target antigens for immunotherapy on primary and metastatic prostate cancers Clin Cancer Res 1998 4 2 295 302
[43]
SS Chang VE Reuter WD Heston NH Bander LS Grauer PB Gaudin Five different anti-prostate-specific membrane antigen (PSMA) antibodies confirm PSMA expression in tumor-associated neovasculature Cancer Res 1999 59 13 3192 8
[44]
RL Sokoloff KC Norton CL Gasior KM Marker LS Grauer A dual-monoclonal sandwich assay for prostate-specific membrane antigen: Levels in tissues, seminal fluid and urine Prostate 2000 43 2 150 157
[45]
Y Kinoshita K Kuratsukuri S Landas K Imaida PM Rovito CY Wang GP Haas Expression of prostate-specific membrane antigen in normal and malignant human tissues World Journal of Surgery 2006 30 4 628 636 DOI:10.1.007/s00268-005-0544-5
[46]
P Mhawech-Fauceglia S Zhang L Terracciano G Sauter A Chadhuri FR Herrmann R Penetrante Prostate-specific membrane antigen (PSMA) protein expression in normal and neoplastic tissues and its sensitivity and specificity in prostate adenocarcinoma: an immunohistochemical study using mutiple tumour tissue microarray technique Histopathology 2007 50 4 472 83 DOI:10.1.111/j.1365-2559.2.007.0.2635.x
[47]
M Rovenska K Hlouchova P Sacha P Mlcochova V Horak J Zamecnik C Barinka J Konvalinka Tissue expression and enzymologic characterization of human prostate specific membrane antigen and its rat and pig orthologs Prostate 2008 68 2 171 82 DOI:10.1.002/pros.20676
[48]
P Wolf N Freudenberg P Buhler K Alt W Schultze-Seemann U Wetterauer U Elsasser-Beile Three conformational antibodies specific for different PSMA epitopes are promising diagnostic and therapeutic tools for prostate cancer Prostate 2010 70 5 562 9 DOI:10.1.002/pros.21090
[49]
P Sacha J Zamecnik C Barinka K Hlouchova A Vicha P Mlcochova I Hilgert T Eckschlager J Konvalinka Expression of glutamate carboxypeptidase II in human brain Neuroscience 2007 144 4 1361 1372 DOI:10.1.016/j.neuroscience.2006.1.0.0.22
[50]
F Dumas JL Gala P Berteau F Brasseur P Eschwege V Paradis B Lacour M Philippe S Loric Molecular expression of PSMA mRNA and protein in primary renal tumors Int J Cancer 1999 80 6 799 803 DOI:10.1.002/(SICI)1097-0215(19990315)80:6<799:: AID-IJC1>3.3.CO;2-L
[51]
UV Berger RE Carter M McKee JT Coyle N-acetylated alpha-linked acidic dipeptidase is expressed by non-myelinating Schwann cells in the peripheral nervous system J Neurocytol 1995 24 2 99 109 DOI:10.1.007/BF01181553
[52]
UV Berger R Luthi-Carter LA Passani S Elkabes I Black C Konradi JT Coyle Glutamate carboxypeptidase II is expressed by astrocytes in the adult rat nervous system Journal of Comparative Neurology 1999 415 1 52 64 DOI:10.1.002/(Sici)1096-9861(19991206)415:1<52:: Aid-Cnne4>3.0.Co;2-K
[53]
VA Carozzi A Canta N Oggioni C Ceresa P Marmiroli J Konvalinka C Zoia M Bossi C Ferrarese G Tredici G Cavaletti Expression and distribution of 'high affinity'glutamate transporters GLT1, GLAST, EAAC1 and of GCPII in the rat peripheral nervous system J Anat 2008 213 5 539 46 DOI:10.1.111/j.1469-7580.2.008.0.0984.x
[54]
YP Rochon JS Horoszewicz AL Boynton EH Holmes RJ Barren 3rd SJ Erickson GM Kenny GP Murphy Western blot assay for prostate-specific membrane antigen in serum of prostate cancer patients Prostate 1994 25 4 219 23 DOI:10.1.002/pros.2990250408
[55]
ML Beckett LH Cazares A Vlahou PF Schellhammer GL Wright Jr Prostate-specific membrane antigen levels in sera from healthy men and patients with benign prostate hyperplasia or prostate cancer Clin Cancer Res 1999 5 12 4034 40
[56]
Z Xiao BL Adam LH Cazares MA Clements JW Davis PF Schellhammer EA Dalmasso GL Wright Jr Quantitation of serum prostate-specific membrane antigen by a novel protein biochip immunoassay discriminates benign from malignant prostate disease Cancer Res 2001 61 16 6029 33
[57]
T Knedlik V Navratil V Vik D Pacik P Sacha J Konvalinka Detection and quantitation of glutamate carboxypeptidase II in human blood Prostate 2014 74 7 768 780 DOI:10.1.002/Pros.22796
[58]
DG Bostwick A Pacelli M Blute P Roche GP Murphy Prostate specific membrane antigen expression in prostatic intraepithelial neoplasia and adenocarcinoma: a study of 184 cases Cancer 1998 82 11 2256 61 DOI:10.1.002/(SICI)1097-0142(19980601)82:11<2256:: AID-CNCR22>3.0.CO;2-S
[59]
S Perner MD Hofer R Kim RB Shah H Li P Moller RE Hautmann JE Gschwend R Kuefer MA Rubin Prostate-specific membrane antigen expression as a predictor of prostate cancer progression Hum Pathol 2007 38 5 696 701 DOI:10.1.016/j.humpath.2006.1.1.0.12
[60]
S Minner C Wittmer M Graefen G Salomon T Steuber A Haese H Huland C Bokemeyer E Yekebas J Dierlamm S Balabanov E Kilic W Wilczak R Simon G Sauter T Schlomm High level PSMA expression is associated with early PSA recurrence in surgically treated prostate cancer Prostate 2011 71 3 281 8 DOI:10.1.002/pros.21241
[61]
A Chaux J Eifler S Karram T Al-Hussain S Faraj M Pomper R Rodriguez GJ Netto Focal positive prostate-specific membrane antigen (PSMA) expression in ganglionic tissues associated with prostate neurovascular bundle: implications for novel intraoperative PSMA-based fluorescent imaging techniques Urol Oncol 2013 31 5 572 5 DOI:10.1.016/j.urolonc.2011.0.4.0.02
[62]
SS Chang VE Reuter WD Heston PB Gaudin Comparison of anti-prostate-specific membrane antigen antibodies and other immunomarkers in metastatic prostate carcinoma Urology 2001 57 6 1179 83 DOI:10.1.016/S0090-4295(01)00983-9
[63]
JL Gala S Loric Y Guiot SR Denmeade A Gady F Brasseur M Heusterspreute P Eschwege P De Nayer P Van Cangh B Tombal Expression of prostate-specific membrane antigen in transitional cell carcinoma of the bladder: prognostic value? Clin Cancer Res 2000 6 10 4049 54
[64]
A Baccala L Sercia J Li W Heston M Zhou Expression of prostate-specific membrane antigen in tumor-associated neovasculature of renal neoplasms Urology 2007 70 2 385 90 DOI:10.1.016/j.urology.2007.0.3.0.25
[65]
Z Lane DE Hansel JI Epstein Immunohistochemical expression of prostatic antigens in adenocarcinoma and villous adenoma of the urinary bladder Am J Surg Pathol 2008 32 9 1322 6 DOI:10.1.097/PAS.0b013e3181656ca0
[66]
MC Haffner IE Kronberger JS Ross CE Sheehan M Zitt G Muhlmann D Ofner B Zelger C Ensinger XJ Yang S Geley R Margreiter NH Bander Prostate-specific membrane antigen expression in the neovasculature of gastric and colorectal cancers Hum Pathol 2009 40 12 1754 61 DOI:10.1.016/j.humpath.2009.0.6.0.03
[67]
MK Samplaski W Heston P Elson C Magi-Galluzzi DE Hansel Folate hydrolase (prostate-specific membrane (corrected) antigen) 1 expression in bladder cancer subtypes and associated tumor neovasculature Mod Pathol 2011 24 11 1521 9 DOI:10.1.038/modpathol.2011.1.12
[68]
C Zeng ZF Ke Z Yang Z Wang SC Yang CQ Luo LT Wang Prostate-specific membrane antigen: a new potential prognostic marker of osteosarcoma Med Oncol 2012 29 3 2234 9 DOI:10.1.007/s12032-011-0089-4
[69]
N Nomura S Pastorino P Jiang G Lambert JR Crawford M Gymnopoulos D Piccioni T Juarez SC Pingle M Makale S Kesari Prostate specific membrane antigen (PSMA) expression in primary gliomas and breast cancer brain metastases Cancer Cell Int 2014 14 1 26 DOI:10.1.186/1475-2867-14-26
[70]
M Abdel-Hadi Y Ismail L Younis Prostate-specific membrane antigen (PSMA) immunoexpression in the neovasculature of colorectal carcinoma in Egyptian patients Pathol Res Pract 2014 210 11 759 63 DOI:10.1.016/j.prp.2014.0.5.0.15
[71]
H Ren H Zhang X Wang J Liu Z Yuan J Hao Prostate-specific membrane antigen as a marker of pancreatic cancer cells Med Oncol 2014 31 3 857 DOI:10.1.007/s12032-014-0857-z
[72]
AG Wernicke S Varma EA Greenwood PJ Christos KS Chao H Liu NH Bander SJ Shin Prostate-specific membrane antigen expression in tumor-associated vasculature of breast cancers APMIS 2014 122 6 482 9 DOI:10.1.111/apm.12195
[73]
HL Wang SS Wang WH Song Y Pan HP Yu TG Si Y Liu XN Cui Z Guo Expression of prostate-specific membrane antigen in lung cancer cells and tumor neovasculature endothelial cells and its clinical significance PLoS One 2015 10 5 e0125924 DOI:10.1.371/journal.pone.0125924
[74]
MJ Crowley T Scognamiglio YF Liu DA Kleiman T Beninato A Aronova H Liu YS Jhanwar A Molina ST Tagawa NH Bander R Zarnegar O Elemento TJ Fahey 3rd Prostate-Specific Membrane Antigen Is a Potential Antiangiogenic Target in Adrenocortical Carcinoma J Clin Endocrinol Metab 2016 101 3 981 7 DOI:10.1.210/jc.2015-4021
[75]
RA Salas Fragomeni JR Menke M Holdhoff C Ferrigno JJ Laterra LB Solnes MS Javadi Z Szabo MG Pomper SP Rowe Prostate-Specific Membrane Antigen-Targeted Imaging With (18F)DCFPyL in High-Grade Gliomas Clin Nucl Med 2017 42 10 e433 e435 DOI:10.1.097/RLU.0000000000001769
[76]
M Kasoha C Unger EF Solomayer RM Bohle C Zaharia F Khreich S Wagenpfeil I Juhasz-Boss Prostate-specific membrane antigen (PSMA) expression in breast cancer and its metastases Clin Exp Metastasis 2017 34 8 479 490 DOI:10.1.007/s10585-018-9878-x
[77]
LH Schmidt B Heitkotter AB Schulze C Schliemann K Steinestel M Trautmann A Marra L Hillejan M Mohr G Evers E Wardelmann K Rahbar D Gorlich G Lenz WE Berdel W Hartmann R Wiewrodt S Huss Prostate specific membrane antigen (PSMA) expression in non-small cell lung cancer PLoS One 2017 12 10 e0186280 DOI:10.1.371/journal.pone.0186280
[78]
B Heitkotter M Trautmann I Grunewald M Bogemann K Rahbar H Gevensleben E Wardelmann W Hartmann K Steinestel S Huss Expression of PSMA in tumor neovasculature of high grade sarcomas including synovial sarcoma, rhabdomyosarcoma, undifferentiated sarcoma and MPNST Oncotarget 2017 8 3 4268 4276 DOI:10.1.8632/oncotarget.13994
[79]
A Bychkov U Vutrapongwatana S Tepmongkol S Keelawat PSMA expression by microvasculature of thyroid tumors - Potential implications for PSMA theranostics Sci Rep 2017 7 1 5202 DOI:10.1.038/s41598-017-05481-z
[80]
M Moore S Panjwani R Mathew M Crowley YF Liu A Aronova B Finnerty R Zarnegar TJ Fahey 3rd T Scognamiglio Well-Differentiated Thyroid Cancer Neovasculature Expresses Prostate-Specific Membrane Antigen-a Possible Novel Therapeutic Target Endocr Pathol 2017 28 4 339 344 DOI:10.1.007/s12022-017-9500-9
[81]
TJW Klein Nulent RJJ van Es GC Krijger R de Bree SM Willems B de Keizer Prostate-specific membrane antigen PET imaging and immunohistochemistry in adenoid cystic carcinoma-a preliminary analysis Eur J Nucl Med Mol Imaging 2017 44 10 1614 1621 DOI:10.1.007/s00259-017-3737-x
[82]
S Spatz Y Tolkach K Jung C Stephan J Busch B Ralla A Rabien G Feldmann P Brossart RA Bundschuh H Ahmadzadehfar M Essler M Toma SC Muller J Ellinger S Hauser G Kristiansen Comprehensive Evaluation of Prostate Specific Membrane Antigen Expression in the Vasculature of Renal Tumors: Implications for Imaging Studies and Prognostic Role J Urol 2018 199 2 370 377 DOI:10.1.016/j.juro.2017.0.8.0.79
[83]
Y Tolkach H Gevensleben R Bundschuh A Koyun D Huber C Kehrer T Hecking MD Keyver-Paik C Kaiser H Ahmadzadehfar M Essler W Kuhn G Kristiansen Prostate-specific membrane antigen in breast cancer: a comprehensive evaluation of expression and a case report of radionuclide therapy Breast Cancer Res Treat 2018 169 3 447 455 DOI:10.1.007/s10549-018-4717-y
[84]
B Heitkotter K Steinestel M Trautmann I Grunewald P Barth H Gevensleben M Bogemann E Wardelmann W Hartmann K Rahbar S Huss Neovascular PSMA expression is a common feature in malignant neoplasms of the thyroid Oncotarget 2018 9 11 9867 9874 DOI:10.1.8632/oncotarget.23984
[85]
M Matsuda E Ishikawa T Yamamoto K Hatano A Joraku Y Iizumi Y Masuda H Nishiyama A Matsumura Potential use of prostate specific membrane antigen (PSMA) for detecting the tumor neovasculature of brain tumors by PET imaging with (89)Zr-Df-IAB2M anti-PSMA minibody J Neurooncol 2018 138 3 581 589 DOI:10.1.007/s11060-018-2825-5
[86]
SS Chang DS O'Keefe DJ Bacich VE Reuter WDW Heston PB Gaudin Prostate-specific membrane antigen is produced in tumor-associated neovasculature Clinical Cancer Research 1999 5 10 2674 2681
[87]
R Rais W Jiang H Zhai KM Wozniak M Stathis KR Hollinger AG Thomas C Rojas JJ Vornov M Marohn X Li BS Slusher FOLH1/GCPII is elevated in IBD patients, and its inhibition ameliorates murine IBD abnormalities JCI Insight 2016 1 12 e88634 DOI:10.1.172/jci.insight.88634
[88]
CL Noble AR Abbas CW Lees J Cornelius K Toy Z Modrusan HF Clark ID Arnott ID Penman J Satsangi L Diehl Characterization of intestinal gene expression profiles in Crohn's disease by genome-wide microarray analysis Inflamm Bowel Dis 2010 16 10 1717 28 DOI:10.1.002/ibd.21263
[89]
T Zhang B Song W Zhu X Xu QQ Gong C Morando T Dassopoulos RD Newberry SR Hunt E Li An ileal Crohn's disease gene signature based on whole human genome expression profiles of disease unaffected ileal mucosal biopsies PLoS One 2012 7 5 e37139 DOI:10.1.371/journal.pone.0037139
[90]
S Ben-Shachar H Yanai L Baram H Elad E Meirovithz A Ofer E Brazowski H Tulchinsky M Pasmanik-Chor I Dotan Gene expression profiles of ileal inflammatory bowel disease correlate with disease phenotype and advance understanding of its immunopathogenesis Inflamm Bowel Dis 2013 19 12 2509 21 DOI:10.1.097/01.MIB.0000437045.2.6036.0.0
[91]
DJ Bacich E Ramadan DS O'Keefe N Bukhari I Wegorzewska O Ojeifo R Olszewski CC Wrenn T Bzdega B Wroblewska WDW Heston JH Neale Deletion of the glutamate carboxypeptidase II gene in mice reveals a second enzyme activity that hydrolyzes N-acetylaspartylglutamate J Neurochem 2002 83 1 20 29 DOI:10.1.046/j.1471-4159.2.002.0.1117.x
[92]
G Tsai KS Dunham U Drager A Grier C Anderson J Collura JT Coyle Early embryonic death of glutamate carboxypeptidase II (NAALADase) homozygous mutants Synapse 2003 50 4 285 92 DOI:10.1.002/syn.10263
[93]
L Han JD Picker LR Schaevitz G Tsai J Feng Z Jiang HC Chu AC Basu J Berger-Sweeney JT Coyle Phenotypic characterization of mice heterozygous for a null mutation of glutamate carboxypeptidase II Synapse 2009 63 8 625 35 DOI:10.1.002/syn.20649
[94]
Y Gao S Xu Z Cui M Zhang Y Lin L Cai Z Wang X Luo Y Zheng Y Wang Q Luo J Jiang JH Neale C Zhong Mice lacking glutamate carboxypeptidase II develop normally, but are less susceptible to traumatic brain injury J Neurochem 2015 134 2 340 53 DOI:10.1.111/jnc.13123
[95]
G Koscielny G Yaikhom V Iyer TF Meehan H Morgan J Atienza-Herrero A Blake CK Chen R Easty A Di Fenza T Fiegel M Grifiths A Horne NA Karp N Kurbatova JC Mason P Matthews DJ Oakley A Qazi J Regnart A Retha LA Santos DJ Sneddon J Warren H Westerberg RJ Wilson DG Melvin D Smedley SD Brown P Flicek WC Skarnes AM Mallon H Parkinson The International Mouse Phenotyping Consortium Web Portal, a unified point of access for knockout mice and related phenotyping data Nucleic Acids Res 2014 42 Release 7.0. D802 9 DOI:10.1.093/nar/gkt977
[96]
B Vorlova F Sedlak P Kasparek K Sramkova M Maly J Zamecnik P Sacha J Konvalinka A novel PSMA/GCPII?deficient mouse model shows enlarged seminal vesicles upon aging Prostate In press 2018 DOI:10.1.002/pros.23717
[97]
DJ Bacich KM Wozniak XCM Lu DS O'Keefe N Callizot WDW Heston BS Slusher Mice lacking glutamate carboxypeptidase II are protected from peripheral neuropathy and ischemic brain injury J Neurochem 2005 95 2 314 323 DOI:10.1.111/j.1471-4159.2.005.0.3361.x
[98]
Y Cao Y Gao S Xu J Bao Y Lin X Luo Y Wang Q Luo J Jiang JH Neale C Zhong Glutamate carboxypeptidase II gene knockout attenuates oxidative stress and cortical apoptosis after traumatic brain injury BMC Neurosci 2016 17 15 DOI:10.1.186/s12868-016-0251-1
[99]
JH Neale T Bzdega B Wroblewska N-acetylaspartylglutamate: The most abundant peptide neurotransmitter in the mammalian central nervous system J Neurochem 2000 75 2 443 452 DOI:10.1.046/j.1471-4159.2.000.0.750443.x
[100]
G Tsai G Forloni MB Robinson BL Stauch JT Coyle Calcium-dependent evoked release of N-(3H)acetylaspartylglutamate from the optic pathway J Neurochem 1988 51 6 1956 9 DOI:10.1.111/j.1471-4159.1.988.tb01186.x
[101]
M Zollinger U Amsler KQ Do P Streit M Cuenod Release of N-acetylaspartylglutamate on depolarization of rat brain slices J Neurochem 1988 51 6 1919 23 DOI:10.1.111/j.1471-4159.1.988.tb01178.x
[102]
B Wroblewska JT Wroblewski OH Saab JH Neale N-Acetylaspartylglutamate Inhibits Forskolin-Stimulated Cyclic-Amp Levels Via a Metabotropic Glutamate-Receptor in Cultured Cerebellar Granule Cells J Neurochem 1993 61 3 943 948 DOI:10.1.111/j.1471-4159.1.993.tb03606.x
[103]
GL Westbrook ML Mayer MA Namboodiri JH Neale High concentrations of N-acetylaspartylglutamate (NAAG) selectively activate NMDA receptors on mouse spinal cord neurons in cell culture J Neurosci 1986 6 11 3385 92
[104]
JH Neale RT Olszewski DY Zuo KJ Janczura CP Profaci KM Lavin JC Madore T Bzdega Advances in understanding the peptide neurotransmitter NAAG and appearance of a new member of the NAAG neuropeptide family J Neurochem 2011 118 4 490 498 DOI:10.1.111/j.1471-4159.2.011.0.7338.x
[105]
P Khacho B Wang R Bergeron The Good and Bad Sides of NAAG Adv Pharmacol 2016 76 311 49 DOI:10.1.016/bs.apha.2016.0.1.0.03
[106]
B Wroblewska JT Wroblewski S Pshenichkin A Surin SE Sullivan JH Neale N-acetylaspartylglutamate selectively activates mGluR3 receptors in transfected cells J Neurochem 1997 69 1 174 181
[107]
B Wroblewska MR Santi JH Neale N-acetylaspartylglutamate activates cyclic AMP-coupled metabotropic glutamate receptors in cerebellar astrocytes Glia 1998 24 2 172 9 DOI:10.1.002/(SICI)1098-1136(199810)24:2<172:: AID-GLIA2>3.0.CO;2-6
[108]
B Wroblewska IN Wegorzewska T Bzdega RT Olszewski JH Neale Differential negative coupling of type 3 metabotropic glutamate receptor to cyclic GMP levels in neurons and astrocytes J Neurochem 2006 96 4 1071 1077 DOI:10.1.111/j.1471-4159.2.005.0.3569.x
[109]
J Zhao E Ramadan M Cappiello B Wroblewska T Bzdega JH Neale NAAG inhibits KCl-induced (H-3)-GABA release via mGluR3, cAMP, PKA and L-type calcium conductance European Journal of Neuroscience 2001 13 2 340 346 DOI:10.1.046/j.1460-9568.2.001.0.1396.x
[110]
C Zhong X Zhao KC Van T Bzdega A Smyth J Zhou AP Kozikowski J Jiang WT O'Connor RF Berman JH Neale BG Lyeth NAAG peptidase inhibitor increases dialysate NAAG and reduces glutamate, aspartate and GABA levels in the dorsal hippocampus following fluid percussion injury in the rat J Neurochem 2006 97 4 1015 25 DOI:10.1.111/j.1471-4159.2.006.0.3786.x
[111]
ERG Sanabria KM Wozniak BS Slusher A Keller GCP II (NAALADase) inhibition suppresses mossy fiber-CA3 synaptic neurotransmission by a presynaptic mechanism Journal of Neurophysiology 2004 91 1 182 193 DOI:10.1.152/jn.00465.2.003
[112]
C Romei M Raiteri L Raiteri Glycine release is regulated by metabotropic glutamate receptors sensitive to mGluR2/3 ligands and activated by N-acetylaspartylglutamate (NAAG) Neuropharmacology 2013 66 311 316 DOI:10.1.016/j.neuropharm.2012.0.5.0.30
[113]
AG Thomas WL Liu JL Olkowski ZC Tang Q Lin XCM Lu BS Slusher Neuroprotection mediated by glutamate carboxypeptidase II (NAALADase) inhibition requires TGF-beta European Journal of Pharmacology 2001 430 1 33 40 DOI:10.1.016/S0014-2999(01)01239-0
[114]
V Bruno G Battaglia G Casabona A Copani F Caciagli F Nicoletti Neuroprotection by glial metabotropic glutamate receptors is mediated by transforming growth factor-beta Journal of Neuroscience 1998 18 23 9594 9600
[115]
A Ruocco O Nicole F Docagne C Ali L Chazalviel S Komesli F Yablonsky S Roussel ET MacKenzie D Vivien A Buisson A transforming growth factor-beta antagonist unmasks the neuroprotective role of this endogenous cytokine in excitotoxic and ischemic brain injury Journal of Cerebral Blood Flow and Metabolism 1999 19 12 1345 1353 DOI:10.1.097/00004647-199912000-00008
[116]
R Zhao LH Matherly ID Goldman Membrane transporters and folate homeostasis: intestinal absorption and transport into systemic compartments and tissues Expert Rev Mol Med 2009 11 e4 DOI:10.1.017/S1462399409000969
[117]
H McNulty K Pentieva LB Bailey Folate Bioavailability In Folate in health and disease 2010 Boca Raton US Taylor & Francis Group
[118]
JW Miller B Caballero Folic Acid In Encyclopedia of Human Nutrition 2013 Waltham MA Academic Press
[119]
CJ Chandler TTY Wang CH Halsted Pteroylpolyglutamate Hydrolase from Human Jejunal Brush-Borders - Purification and Characterization J Biol Chem 1986 261 2 928 933
[120]
A Qiu M Jansen A Sakaris SH Min S Chattopadhyay E Tsai C Sandoval R Zhao MH Akabas ID Goldman Identification of an intestinal folate transporter and the molecular basis for hereditary folate malabsorption Cell 2006 127 5 917 28 DOI:10.1.016/j.cell.2006.0.9.0.41
[121]
A Qiu SH Min M Jansen U Malhotra E Tsai DC Cabelof LH Matherly R Zhao MH Akabas ID Goldman Rodent intestinal folate transporters (SLC46A1): secondary structure, functional properties, and response to dietary folate restriction Am J Physiol Cell Physiol 2007 293 5 C1669 78 DOI:10.1.152/ajpcell.00202.2.007
[122]
AM Devlin EH Ling JM Peerson S Fernando R Clarke AD Smith CH Halsted Glutamate carboxypeptidase II: a polymorphism associated with lower levels of serum folate and hyperhomocysteinemia Human Molecular Genetics 2000 9 19 2837 2844 DOI:10.1.093/hmg/9.1.9.2.837
[123]
KJA Lievers LAJ Kluijtmans GHJ Boers P Verhoef M den Heijer FJM Trijbels HJ Blom Influence of a glutamate carboxypeptidase II (GCPII) polymorphism (1561C ->T) on plasma homocysteine, folate and vitamin B-12 levels and its relationship to cardiovascular disease risk Atherosclerosis 2002 164 2 269 273 DOI:10.1.016/S0021-9150(02)00065-5
[124]
C Vargas-Martinez JM Ordovas PW Wilson J Selhub The glutamate carboxypeptidase gene II (C>T) polymorphism does not affect folate status in the Framingham Offspring cohort J Nutr 2002 132 6 1176 9
[125]
I Morin AM Devlin D Leclerc N Sabbaghian CH Halsted R Finnell R Rozen Evaluation of genetic variants in the reduced folate carrier and in glutamate carboxypeptidase II for spina bifida risk Molecular Genetics and Metabolism 2003 79 3 197 200 DOI: S1096719203000866
[126]
LA Afman FJM Trijbels HJ Blom The H475Y polymorphism in the glutamate carboxypeptidase II gene increases plasma folate without affecting the risk for neural tube defects in humans Journal of Nutrition 2003 133 1 75 77
[127]
A Melse-Boonstra KJA Lievers HJ Blom P Verhoef Bioavailability of polyglutamyl folic acid relative to that of monoglutamyl folic acid in subjects with different genotypes of the glutamate carboxypeptidase II gene American Journal of Clinical Nutrition 2004 80 3 700 704
[128]
J Chen C Kyte M Valcin W Chan JG Wetmur J Selhub DJ Hunter J Ma Polymorphisms in the one-carbon metabolic pathway, plasma folate levels and colorectal cancer in a prospective study Int J Cancer 2004 110 4 617 20 DOI:10.1.002/ijc.20148
[129]
AM Devlin R Clarke J Birks JG Evans CH Halsted Interactions among polymorphisms in folate-metabolizing genes and serum total homocysteine concentrations in a healthy elderly population Am J Clin Nutr 2006 83 3 708 13 DOI:83/3/708
[130]
CH Halsted DH Wong JM Peerson CH Warden H Refsum AD Smith OK Nygard PM Ueland SE Vollset GS Tell Relations of glutamate carboxypeptidase II (GCPII) polymorphisms to folate and homocysteine concentrations and to scores of cognition, anxiety, and depression in a homogeneous Norwegian population: the Hordaland Homocysteine Study Am J Clin Nutr 2007 86 2 514 21 DOI:86/2/514
[131]
L DeVos A Chanson ZH Liu ED Ciappio LD Parnell JB Mason KL Tucker JW Crott Associations between single nucleotide polymorphisms in folate uptake and metabolizing genes with blood folate, homocysteine, and DNA uracil concentrations American Journal of Clinical Nutrition 2008 88 4 1149 1158
[132]
V Eklof B Van Guelpen J Hultdin I Johansson G Hallmans R Palmqvist The reduced folate carrier (RFC1) 80G >A and folate hydrolase 1 (FOLH1) 1561C >T polymorphisms and the risk of colorectal cancer: a nested case-referent study Scand J Clin Lab Invest 2008 68 5 393 401 DOI:10.1.080/00365510701805431
[133]
J Guo H Xie J Wang H Zhao F Wang C Liu L Wang X Lu Y Bao J Zou G Wang B Niu T Zhang The maternal folate hydrolase gene polymorphism is associated with neural tube defects in a high-risk Chinese population Genes Nutr 2013 8 2 191 7 DOI:10.1.007/s12263-012-0309-3
[134]
JL Roffman JS Lamberti E Achtyes EA Macklin GC Galendez LH Raeke NJ Silverstein JW Smoller M Hill DC Goff Randomized multicenter investigation of folate plus vitamin B12 supplementation in schizophrenia JAMA Psychiatry 2013 70 5 481 9 DOI:10.1.001/jamapsychiatry.2013.9.00
[135]
D Cummings KF Dowling NJ Silverstein AS Tanner H Eryilmaz JW Smoller JL Roffman A Cross-Sectional Study of Dietary and Genetic Predictors of Blood Folate Levels in Healthy Young Adults Nutrients 2017 9 9 DOI:10.3.390/nu9090994
[136]
JL Roffman DG Brohawn AZ Nitenson EA Macklin JW Smoller DC Goff Genetic variation throughout the folate metabolic pathway influences negative symptom severity in schizophrenia Schizophr Bull 2013 39 2 330 8 DOI:10.1.093/schbul/sbr150
[137]
AR Vieira D Trembath DC Vandyke JC Murray S Marker G Lerner E Bonner M Speer Studies with His475Tyr glutamate carboxipeptidase II polymorphism and neural tube defects American Journal of Medical Genetics 2002 111 2 218 219 DOI:10.1.002/Ajmg.10568
[138]
H Xie J Guo J Wang F Wang H Zhao C Liu L Wang X Lu L Wu Y Bao J Zou T Zhang B Niu Glutamate carboxypeptidase II gene polymorphisms and neural tube defects in a high-risk Chinese population Metab Brain Dis 2012 27 1 59 65 DOI:10.1.007/s11011-011-9272-8
[139]
CL Relton CS Wilding MS Pearce AJ Laffling PA Jonas SA Lynch EJ Tawn J Burn Gene-gene interaction in folate-related genes and risk of neural tube defects in a UK population J Med Genet 2004 41 4 256 60 DOI:10.1.136/jmg.2003.0.10694
[140]
TO Findley JC Tenpenny MR O'Byrne AC Morrison JE Hixson H Northrup KS Au Mutations in folate transporter genes and risk for human myelomeningocele Am J Med Genet A 2017 173 11 2973 2984 DOI:10.1.002/ajmg.a.38472
[141]
A Hazra K Wu P Kraft CS Fuchs EL Giovannucci DJ Hunter Twenty-four non-synonymous polymorphisms in the one-carbon metabolic pathway and risk of colorectal adenoma in the Nurses'Health Study Carcinogenesis 2007 28 7 1510 9 DOI:10.1.093/carcin/bgm062
[142]
T Gotze C Rocken FW Rohl T Wex J Hoffmann S Westphal P Malfertheiner MP Ebert J Dierkes Gene polymorphisms of folate metabolizing enzymes and the risk of gastric cancer Cancer Lett 2007 251 2 228 36 DOI:10.1.016/j.canlet.2006.1.1.0.21
[143]
H Liu G Jin H Wang W Wu Y Liu J Qian W Fan H Ma R Miao Z Hu W Sun Y Wang L Jin Q Wei H Shen W Huang D Lu Association of polymorphisms in one-carbon metabolizing genes and lung cancer risk: a case-control study in Chinese population Lung Cancer 2008 61 1 21 9 DOI:10.1.016/j.lungcan.2007.1.2.0.01
[144]
MM Mir Combined impact of polymorphism of folate metabolism genes;glutamate carboxypeptidase, methylene tetrahydrofolate reductase and methionine synthase reductase on breast cancer susceptibility in Kashmiri women Clinical Chemistry 2008 54 6 A126 A126
[145]
CM Lawrence S Ray M Babyonyshev R Galluser DW Borhani SC Harrison Crystal structure of the ectodomain of human transferrin receptor Science 1999 286 5440 779 82 DOI:10.1.126/science.286.5.440.7.79
[146]
H Liu AK Rajasekaran P Moy Y Xia S Kim V Navarro R Rahmati NH Bander Constitutive and antibody-induced internalization of prostate-specific membrane antigen Cancer Res 1998 58 18 4055 60
[147]
SA Rajasekaran G Anilkumar E Oshima JU Bowie H Liu W Heston NH Bander AK Rajasekaran A novel cytoplasmic tail MXXXL motif mediates the internalization of prostate-specific membrane antigen Mol Biol Cell 2003 14 12 4835 45 DOI:10.1.091/mbce02-11-0731
[148]
V Yao CE Berkman JK Choi DS O'Keefe DJ Bacich Expression of Prostate-Specific Membrane Antigen (PSMA), Increases Cell Folate Uptake and Proliferation and Suggests a Novel Role for PSMA in the Uptake of the Non-Polyglutamated Folate, Folic Acid Prostate 2010 70 3 305 316 DOI:10.1.002/Pros.21065
[149]
KS Au TO Findley H Northrup Finding the genetic mechanisms of folate deficiency and neural tube defects-Leaving no stone unturned Am J Med Genet A 2017 173 11 3042 3057 DOI:10.1.002/ajmg.a.38478
[150]
V Yao A Parwani C Maier WD Heston DJ Bacich Moderate expression of prostate-specific membrane antigen, a tissue differentiation antigen and folate hydrolase, facilitates prostate carcinogenesis Cancer Res 2008 68 21 9070 7 DOI:10.1.158/0008-5472.CAN-08-2328
[151]
A Ghosh X Wang E Klein WD Heston Novel role of prostate-specific membrane antigen in suppressing prostate cancer invasiveness Cancer Res 2005 65 3 727 31
[152]
M Miyake Y Kakimoto M Sorimachi Isolation and Identification of Beta-Citryl-L-Glutamic Acid from Newborn Rat-Brain Biochimica Et Biophysica Acta 1978 544 3 656 666 DOI:10.1.016/0304-4165(78)90340-9
[153]
M Miyake Y Kakimoto Developmental-Changes of N-Acetyl-L-Aspartic Acid, N-Acetyl-Alpha-Aspartylglutamic Acid and Beta-Citryl-L-Glutamic Acid in Different Brain-Regions and Spinal Cords of Rat and Guinea Pig J Neurochem 1981 37 4 1064 1067 DOI:10.1.111/j.1471-4159.1.981.tb04500.x
[154]
M Miyake S Kume Y Kakimoto Correlation of the level of beta-citryl-L-glutamic acid with spermatogenesis in rat testes Biochim Biophys Acta 1982 719 3 495 500 DOI: 0304-4165(82)90238-0
[155]
M Hamada-Kanazawa M Kouda A Odani K Matsuyama K Kanazawa T Hasegawa M Narahara M Miyake beta-Citryl-L-glutamate Is an Endogenous Iron Chelator That Occurs Naturally in the Developing Brain Biological &Pharmaceutical Bulletin 2010 33 5 729 737 DOI:10.1.248/bpb.33.7.29
[156]
M Narahara M Hamada-Kanazawa M Kouda A Odani M Miyake Superoxide Scavenging and Xanthine Oxidase Inhibiting Activities of Copper S-Citryl-L-glutamate Complex Biological &Pharmaceutical Bulletin 2010 33 12 1938 1943 DOI:10.1.248/Bpb.33.1.938
[157]
M Hamada-Kanazawa M Narahara M Takano KS Min K Tanaka M Miyake beta-Citryl-L-glutamate Acts as an Iron Carrier to Activate Aconitase Activity Biological &Pharmaceutical Bulletin 2011 34 9 1455 1464 DOI:10.1.248/bpb.34.1.455
[158]
T Wustemann U Haberkorn J Babich W Mier Targeting prostate cancer: Prostate-specific membrane antigen based diagnosis and therapy Med Res Rev 2018 DOI:10.1.002/med.21508
[159]
PF Jackson DC Cole BS Slusher SL Stetz LE Ross BA Donzanti DA Trainor Design, synthesis, and biological activity of a potent inhibitor of the neuropeptidase N-acetylated alpha-linked acidic dipeptidase J Med Chem 1996 39 2 619 22 DOI:10.1.021/jm950801q
[160]
AP Kozikowski F Nan P Conti J Zhang E Ramadan T Bzdega B Wroblewska JH Neale S Pshenichkin JT Wroblewski Design of remarkably simple, yet potent urea-based inhibitors of glutamate carboxypeptidase II (NAALADase) J Med Chem 2001 44 3 298 301 DOI:10.1.021/jm000406m
[161]
C Barinka K Hlouchova M Rovenska P Majer M Dauter N Hin YS Ko T Tsukamoto BS Slusher J Konvalinka J Lubkowski Structural basis of interactions between human glutamate carboxypeptidase II and its substrate analogs Journal of Molecular Biology 2008 376 5 1438 1450 DOI:10.1.016/j.jmb.2007.1.2.0.66
[162]
P Majer PF Jackson G Delahanty BS Grella YS Ko W Li Q Liu KM Maclin J Polakova KA Shaffer D Stoermer D Vitharana EY Wang A Zakrzewski C Rojas BS Slusher KM Wozniak E Burak T Limsakun T Tsukamoto Synthesis and biological evaluation of thiol-based inhibitors of glutamate carboxypeptidase II: discovery of an orally active GCP II inhibitor J Med Chem 2003 46 10 1989 96 DOI:10.1.021/jm020515w
[163]
Y Chen CA Foss Y Byun S Nimmagadda M Pullambhatla JJ Fox M Castanares SE Lupold JW Babich RC Mease MG Pomper Radiohalogenated prostate-specific membrane antigen (PSMA)-based ureas as imaging agents for prostate cancer J Med Chem 2008 51 24 7933 43 DOI:10.1.021/jm801055h
[164]
T Yamamoto S Hirasawa B Wroblewska E Grajkowska J Zhou A Kozikowski J Wroblewski JH Neale Antinociceptive effects of N-acetylaspartylglutamate (NAAG) peptidase inhibitors ZJ-11, ZJ-17 and ZJ-43 in the rat formalin test and in the rat neuropathic pain model Eur J Neurosci 2004 20 2 483 94 DOI:10.1.111/j.1460-9568.2.004.0.3504.x
[165]
J Zhou JH Neale MG Pomper AP Kozikowski NAAG peptidase inhibitors and their potential for diagnosis and therapy Nat Rev Drug Discov 2005 4 12 1015 26 DOI:10.1.038/nrd1903
[166]
DV Ferraris K Shukla T Tsukamoto Structure-activity relationships of glutamate carboxypeptidase II (GCPII) inhibitors Curr Med Chem 2012 19 9 1282 94
[167]
P Majer A Jancarik M Krecmerova T Tichy L Tenora K Wozniak Y Wu E Pommier D Ferraris R Rais BS Slusher Discovery of Orally Available Prodrugs of the Glutamate Carboxypeptidase II (GCPII) Inhibitor 2-Phosphonomethylpentanedioic Acid (2-PMPA) J Med Chem 2016 59 6 2810 9 DOI:10.1.021/acs.jmedchem.6b00062
[168]
G Alquicer D Sedlak Y Byun J Pavlicek M Stathis C Rojas B Slusher MG Pomper P Bartunek C Barinka Development of a High-Throughput Fluorescence Polarization Assay to Identify Novel Ligands of Glutamate Carboxypeptidase II Journal of Biomolecular Screening 2012 17 8 1030 1040 DOI:10.1.177/1087057112451924
[169]
V Navratil J Schimer J Tykvart T Knedlik V Vik P Majer J Konvalinka P Sacha DNA-linked Inhibitor Antibody Assay (DIANA) for sensitive and selective enzyme detection and inhibitor screening Nucleic Acids Res 2017 45 2 e10 DOI:10.1.093/nar/gkw853
[170]
J Tykvart J Schimer A Jancarik J Barinkova V Navratil J Starkova K Sramkova J Konvalinka P Majer P Sacha Design of Highly Potent Urea-Based, Exosite-Binding Inhibitors Selective for Glutamate Carboxypeptidase II J Med Chem 2015 58 10 4357 63 DOI:10.1.021/acs.jmedchem.5b00278
[171]
JH Neale RT Olszewski LM Gehl B Wroblewska T Bzdega The neurotransmitter N-acetylaspartylglutamate in models of pain, ALS, diabetic neuropathy, CNS injury and schizophrenia Trends in Pharmacological Sciences 2005 26 9 477 484 DOI:10.1.016/j.tips.2005.0.7.0.04
[172]
BS Slusher JJ Vornov AG Thomas PD Hurn I Harukuni A Bhardwaj RJ Traystman MB Robinson P Britton XC Lu FC Tortella KM Wozniak M Yudkoff BM Potter PF Jackson Selective inhibition of NAALADase, which converts NAAG to glutamate, reduces ischemic brain injury Nature Medicine 1999 5 12 1396 402 DOI:10.1.038/70971
[173]
FC Tortella Y Lin H Ved BS Slusher JR Dave Neuroprotection produced by the NAALADase inhibitor 2-PMPA in rat cerebellar neurons Eur J Pharmacol 2000 402 1-2 31 7 DOI:10.1.016/S0014-2999(00)00519-7
[174]
C Zhong X Zhao J Sarva A Kozikowski JH Neale BG Lyeth NAAG peptidase inhibitor reduces acute neuronal degeneration and astrocyte damage following lateral fluid percussion TBI in rats J Neurotrauma 2005 22 2 266 76 DOI:10.1.089/neu.2005.2.2.2.66
[175]
JF Feng GG Gurkoff KC Van M Song DA Lowe J Zhou BG Lyeth NAAG peptidase inhibitor reduces cellular damage in a model of TBI with secondary hypoxia Brain Res 2012 1469 144 52 DOI:10.1.016/j.brainres.2012.0.6.0.21
[176]
GD Ghadge BS Slusher A Bodner MD Canto K Wozniak AG Thomas C Rojas T Tsukamoto P Majer RJ Miller AL Monti RP Roos Glutamate carboxypeptidase II inhibition protects motor neurons from death in familial amyotrophic lateral sclerosis models Proc Natl Acad Sci U S A 2003 100 16 9554 9 DOI:10.1.073/pnas.1530168100
[177]
T Yamamoto N Nozaki-Taguchi Y Sakashita T Inagaki Inhibition of spinal N-acetylated-alpha-linked acidic dipeptidase produces an antinociceptive effect in the rat formalin test Neuroscience 2001 102 2 473 9 DOI:10.1.016/S0306-4522(00)00502-9
[178]
KJ Carpenter S Sen EA Matthews SL Flatters KM Wozniak BS Slusher AH Dickenson Effects of GCP-II inhibition on responses of dorsal horn neurones after inflammation and neuropathy: an electrophysiological study in the rat Neuropeptides 2003 37 5 298 306
[179]
RT Olszewski N Bukhari J Zhou AP Kozikowski JT Wroblewski S Shamimi-Noori B Wroblewska T Bzdega S Vicini FB Barton JH Neale NAAG peptidase inhibition reduces locomotor activity and some stereotypes in the PCP model of schizophrenia via group II mGluR J Neurochem 2004 89 4 876 885 DOI:10.1.111/j.1471-4159.2.004.0.2358.x
[180]
RT Olszewski KJ Janczura SR Ball JC Madore KM Lavin JCM Lee MJ Lee EK Der TJ Hark PR Farago CP Profaci T Bzdega JH Neale NAAG peptidase inhibitors block cognitive deficit induced by MK-801 and motor activation induced by d-amphetamine in animal models of schizophrenia Translational Psychiatry 2012 2 DOI:10.1.038/tp.2012.6.8
[181]
KA Rahn CC Watkins J Alt R Rais M Stathis I Grishkan CM Crainiceau MG Pomper C Rojas MV Pletnikov PA Calabresi J Brandt PB Barker BS Slusher AI Kaplin Inhibition of glutamate carboxypeptidase II (GCPII) activity as a treatment for cognitive impairment in multiple sclerosis Proc Natl Acad Sci U S A 2012 109 49 20101 6 DOI:10.1.073/pnas.1209934109
[182]
D Ha SJ Bing G Ahn J Kim J Cho A Kim KH Herath HS Yu SA Jo IH Cho Y Jee Blocking glutamate carboxypeptidase II inhibits glutamate excitotoxicity and regulates immune responses in experimental autoimmune encephalomyelitis FEBS J 2016 283 18 3438 56 DOI:10.1.111/febs.13816
[183]
T Knedlik B Vorlova V Navratil J Tykvart F Sedlak S Vaculin M Franek P Sacha J Konvalinka Mouse glutamate carboxypeptidase II (GCPII) has a similar enzyme activity and inhibition profile but a different tissue distribution to human GCPII FEBS Open Bio 2017 7 9 1362 1378 DOI:10.1.002/2211-5463.1.2276
[184]
CG Parsons W Danysz G Quack Glutamate in CNS disorders as a target for drug development: an update Drug News Perspect 1998 11 9 523 69 DOI:10.1.358/dnp.1998.1.1.9.8.6.3689
[185]
JP van der Post SJ de Visser ML de Kam M Woelfler DC Hilt J Vornov ES Burak E Bortey BS Slusher T Limsakun AF Cohen JM van Gerven The central nervous system effects, pharmacokinetics and safety of the NAALADase-inhibitor GPI5693 Br J Clin Pharmacol 2005 60 2 128 36 DOI:10.1.111/j.1365-2125.2.005.0.2396.x
[186]
IA Jaffe Adverse effects profile of sulfhydryl compounds in man Am J Med 1986 80 3 471 6 DOI:10.1.016/0002-9343(86)90722-9
[187]
DV Ferraris P Majer C Ni CE Slusher R Rais Y Wu KM Wozniak J Alt C Rojas BS Slusher T Tsukamoto delta-Thiolactones as prodrugs of thiol-based glutamate carboxypeptidase II (GCPII) inhibitors J Med Chem 2014 57 1 243 7 DOI:10.1.021/jm401703a
[188]
M Nedelcovych RP Dash L Tenora SC Zimmermann AJ Gadiano C Garrett J Alt KR Hollinger E Pommier A Jancarik C Rojas AG Thomas Y Wu K Wozniak P Majer BS Slusher R Rais Enhanced Brain Delivery of (2-(phosphonomethyl)pentanedioic acid) following Intranasal Administration of its gamma-substituted Ester Prodrugs Mol Pharm 2017 DOI:10.1.021/acs.molpharmaceut.7b00231
[189]
R Rais J Vavra T Tichy RP Dash AJ Gadiano L Tenora L Monincova C Barinka J Alt SC Zimmermann CE Slusher Y Wu K Wozniak P Majer T Tsukamoto BS Slusher Discovery of a para-acetoxy-benzyl ester prodrug of a hydroxamate-based Glutamate Carboxypeptidase II inhibitor as oral therapy for neuropathic pain J Med Chem 2017 DOI:10.1.021/acs.jmedchem.7b00825
[190]
AA Date R Rais T Babu J Ortiz P Kanvinde AG Thomas SC Zimmermann AJ Gadiano G Halpert BS Slusher LM Ensign Local enema treatment to inhibit FOLH1/GCPII as a novel therapy for inflammatory bowel disease J Control Release 2017 DOI:10.1.016/j.jconrel.2017.0.1.0.36
[191]
RA Salas Fragomeni T Amir S Sheikhbahaei SC Harvey MS Javadi LB Solnes AP Kiess ME Allaf MG Pomper MA Gorin SP Rowe Imaging of Nonprostate Cancers Using PSMA-Targeted Radiotracers: Rationale, Current State of the Field, and a Call to Arms J Nucl Med 2018 59 6 871 877 DOI:10.2.967/jnumed.117.2.03570
[192]
L Fass Imaging and cancer: a review Mol Oncol 2008 2 2 115 52 DOI:10.1.016/j.molonc.2008.0.4.0.01
[193]
R Weissleder MJ Pittet Imaging in the era of molecular oncology Nature 2008 452 7187 580 9 DOI:10.1.038/naturne06917
[194]
D Kahn RD Williams DW Seldin JA Libertino M Hirschhorn R Dreicer GJ Weiner D Bushnell J Gulfo Radioimmunoscintigraphy with 111indium labeled CYT-356 for the detection of occult prostate cancer recurrence J Urol 1994 152 5 Pt 1 1490 5 DOI:10.1.016/S0022-5347(17)32453-9
[195]
MK Haseman NL Reed SA Rosenthal Monoclonal antibody imaging of occult prostate cancer in patients with elevated prostate-specific antigen. Positron emission tomography and biopsy correlation Clin Nucl Med 1996 21 9 704 13 DOI:10.1.097/00003072-199609000-00007
[196]
N Pandit-Taskar JA O'Donoghue JC Durack SK Lyashchenko SM Cheal V Beylergil RA Lefkowitz JA Carrasquillo DF Martinez AM Fung SB Solomon M Gonen G Heller M Loda DM Nanus ST Tagawa JL Feldman JR Osborne JS Lewis VE Reuter WA Weber NH Bander HI Scher SM Larson MJ Morris A Phase I/II Study for Analytic Validation of 89Zr-J591 ImmunoPET as a Molecular Imaging Agent for Metastatic Prostate Cancer Clin Cancer Res 2015 21 23 5277 85 DOI:10.1.158/1078-0432.CCR-15-0552
[197]
ST Tagawa MI Milowsky M Morris S Vallabhajosula P Christos NH Akhtar J Osborne SJ Goldsmith S Larson NP Taskar HI Scher NH Bander DM Nanus Phase II study of Lutetium-177-labeled anti-prostate-specific membrane antigen monoclonal antibody J591 for metastatic castration-resistant prostate cancer Clin Cancer Res 2013 19 18 5182 91 DOI:10.1.158/1078-0432.CCR-13-0231
[198]
M Eder M Schafer U Bauder-Wust WE Hull C Wangler W Mier U Haberkorn M Eisenhut 68Ga-complex lipophilicity and the targeting property of a urea-based PSMA inhibitor for PET imaging Bioconjug Chem 2012 23 4 688 97 DOI:10.1.021/bc200279b
[199]
Y Chen M Pullambhatla CA Foss Y Byun S Nimmagadda S Senthamizhchelvan G Sgouros RC Mease MG Pomper 2-(3-{1-Carboxy-5-((6-(18F)fluoro-pyridine-3-carbonyl)-amino)-pentyl}-ureido)-pen tanedioic acid, (18F)DCFPyL, a PSMA-based PET imaging agent for prostate cancer Clin Cancer Res 2011 17 24 7645 53 DOI:10.1.158/1078-0432.CCR-11-1357
[200]
A Afshar-Oromieh E Avtzi FL Giesel T Holland-Letz HG Linhart M Eder M Eisenhut S Boxler BA Hadaschik C Kratochwil W Weichert K Kopka J Debus U Haberkorn The diagnostic value of PET/CT imaging with the (68)Ga-labelled PSMA ligand HBED-CC in the diagnosis of recurrent prostate cancer Eur J Nucl Med Mol Imaging 2015 42 2 197 209 DOI:10.1.007/s00259-014-2949-6
[201]
A Afshar-Oromieh CM Zechmann A Malcher M Eder M Eisenhut HG Linhart T Holland-Letz BA Hadaschik FL Giesel J Debus U Haberkorn Comparison of PET imaging with a (68)Ga-labelled PSMA ligand and (18)F-choline-based PET/CT for the diagnosis of recurrent prostate cancer Eur J Nucl Med Mol Imaging 2014 41 1 11 20 DOI:10.1.007/s00259-013-2525-5
[202]
A Afshar-Oromieh A Malcher M Eder M Eisenhut HG Linhart BA Hadaschik T Holland-Letz FL Giesel C Kratochwil S Haufe U Haberkorn CM Zechmann PET imaging with a (68Ga)gallium-labelled PSMA ligand for the diagnosis of prostate cancer: biodistribution in humans and first evaluation of tumour lesions Eur J Nucl Med Mol Imaging 2013 40 4 486 95 DOI:10.1.007/s00259-012-2298-2
[203]
M Dietlein C Kobe G Kuhnert S Stockter T Fischer K Schomacker M Schmidt F Dietlein BD Zlatopolskiy P Krapf R Richarz S Neubauer A Drzezga B Neumaier Comparison of ((18)F)DCFPyL and ((68)Ga)Ga-PSMA-HBED-CC for PSMA-PET Imaging in Patients with Relapsed Prostate Cancer Mol Imaging Biol 2015 17 4 575 84 DOI:10.1.007/s11307-015-0866-0
[204]
Z Szabo E Mena SP Rowe D Plyku R Nidal MA Eisenberger ES Antonarakis H Fan RF Dannals Y Chen RC Mease M Vranesic A Bhatnagar G Sgouros SY Cho MG Pomper Initial Evaluation of ((18)F)DCFPyL for Prostate-Specific Membrane Antigen (PSMA)-Targeted PET Imaging of Prostate Cancer Mol Imaging Biol 2015 17 4 565 74 DOI:10.1.007/s11307-015-0850-8
[205]
NP Lenzo D Meyrick JH Turner Review of Gallium-68 PSMA PET/CT Imaging in the Management of Prostate Cancer Diagnostics (Basel) 2018 8 1 DOI:10.3.390/diagnostics8010016
[206]
C Liu T Liu N Zhang Y Liu N Li P Du Y Yang M Liu K Gong X Yang H Zhu K Yan Z Yang (68)Ga-PSMA-617 PET/CT: a promising new technique for predicting risk stratification and metastatic risk of prostate cancer patients Eur J Nucl Med Mol Imaging 2018 DOI:10.1.007/s00259-018-4037-9
[207]
S Robu A Schmidt M Eiber M Schottelius T Gunther B Hooshyar Yousefi M Schwaiger HJ Wester Synthesis and preclinical evaluation of novel (18)F-labeled Glu-urea-Glu-based PSMA inhibitors for prostate cancer imaging: a comparison with (18)F-DCFPyl and (18)F-PSMA-1007 EJNMMI Res 2018 8 1 30 DOI:10.1.186/s13550-018-0382-8
[208]
SH Moon MK Hong YJ Kim YS Lee DS Lee JK Chung JM Jeong Development of a Ga-68 labeled PET tracer with short linker for prostate-specific membrane antigen (PSMA) targeting Bioorg Med Chem 2018 26 9 2501 2507 DOI:10.1.016/j.bmc.2018.0.4.0.14
[209]
D Matsuoka H Watanabe Y Shimizu H Kimura Y Yagi R Kawai M Ono H Saji Structure-activity relationships of succinimidyl-Cys-C(O)-Glu derivatives with different near-infrared fluorophores as optical imaging probes for prostate-specific membrane antigen Bioorg Med Chem 2018 26 9 2291 2301 DOI:10.1.016/j.bmc.2018.0.3.0.15
[210]
CA Umbricht M Benesova R Schibli C Muller Preclinical Development of Novel PSMA-Targeting Radioligands: Modulation of Albumin-Binding Properties To Improve Prostate Cancer Therapy Mol Pharm 2018 15 6 2297 2306 DOI:10.1.021/acs.molpharmaceut.8b00152
[211]
E Eppard A de la Fuente M Benesova A Khawar RA Bundschuh FC Gartner B Kreppel K Kopka M Essler F Rosch Clinical Translation and First In-Human Use of ((44)Sc)Sc-PSMA-617 for PET Imaging of Metastasized Castrate-Resistant Prostate Cancer Theranostics 2017 7 18 4359 4369 DOI:10.7.150/thno.20586
[212]
J Kelly A Amor-Coarasa S Ponnala A Nikolopoulou C Williams Jr D Schlyer Y Zhao D Kim JW Babich Trifunctional PSMA-targeting constructs for prostate cancer with unprecedented localization to LNCaP tumors Eur J Nucl Med Mol Imaging 2018 DOI:10.1.007/s00259-018-4004-5
[213]
N Deb M Goris K Trisler S Fowler J Saal S Ning M Becker C Marquez S Knox Treatment of hormone-refractory prostate cancer with 90Y-CYT-356 monoclonal antibody Clin Cancer Res 1996 2 8 1289 97
[214]
D Kahn JC Austin RT Maguire SJ Miller J Gerstbrein RD Williams A phase II study of (90Y) yttrium-capromab pendetide in the treatment of men with prostate cancer recurrence following radical prostatectomy Cancer Biother Radiopharm 1999 14 2 99 111 DOI:10.1.089/cbr.1999.1.4.9.9
[215]
PM Smith-Jones S Vallabahajosula SJ Goldsmith V Navarro CJ Hunter D Bastidas NH Bander In vitro characterization of radiolabeled monoclonal antibodies specific for the extracellular domain of prostate-specific membrane antigen. Cancer Res 2000 60 18 5237 43 DOI:10.1.021/jm800765e
[216]
ST Tagawa H Beltran S Vallabhajosula SJ Goldsmith J Osborne D Matulich K Petrillo S Parmar DM Nanus NH Bander Anti-prostate-specific membrane antigen-based radioimmunotherapy for prostate cancer Cancer 2010 116 4 Suppl 1075 83 DOI:10.1.002/cncr.24795
[217]
NH Bander MI Milowsky DM Nanus L Kostakoglu S Vallabhajosula SJ Goldsmith Phase I trial of 177lutetium-labeled J591, a monoclonal antibody to prostate-specific membrane antigen, in patients with androgen-independent prostate cancer J Clin Oncol 2005 23 21 4591 601 DOI:10.1.200/JCO.2005.0.5.1.60
[218]
Z Novakova CA Foss BT Copeland V Morath P Baranova B Havlinova A Skerra MG Pomper C Barinka Novel Monoclonal Antibodies Recognizing Human Prostate-Specific Membrane Antigen (PSMA) as Research and Theranostic Tools Prostate 2017 77 7 749 764 DOI:10.1.002/pros.23311
[219]
M Benesova M Schafer U Bauder-Wust A Afshar-Oromieh C Kratochwil W Mier U Haberkorn K Kopka M Eder Preclinical Evaluation of a Tailor-Made DOTA-Conjugated PSMA Inhibitor with Optimized Linker Moiety for Imaging and Endoradiotherapy of Prostate Cancer J Nucl Med 2015 56 6 914 20 DOI:10.2.967/jnumed.114.1.47413
[220]
C Kratochwil FL Giesel M Stefanova M Benesova M Bronzel A Afshar-Oromieh W Mier M Eder K Kopka U Haberkorn PSMA-Targeted Radionuclide Therapy of Metastatic Castration-Resistant Prostate Cancer with 177Lu-Labeled PSMA-617 J Nucl Med 2016 57 8 1170 6 DOI:10.2.967/jnumed.115.1.71397
[221]
RP Baum HR Kulkarni C Schuchardt A Singh M Wirtz S Wiessalla M Schottelius D Mueller I Klette HJ Wester 177Lu-Labeled Prostate-Specific Membrane Antigen Radioligand Therapy of Metastatic Castration-Resistant Prostate Cancer: Safety and Efficacy J Nucl Med 2016 57 7 1006 13 DOI:10.2.967/jnumed.115.1.68443
[222]
K Rahbar H Ahmadzadehfar C Kratochwil U Haberkorn M Schafers M Essler RP Baum HR Kulkarni M Schmidt A Drzezga P Bartenstein A Pfestroff M Luster U Lutzen M Marx V Prasad W Brenner A Heinzel FM Mottaghy J Ruf PT Meyer M Heuschkel M Eveslage M Bogemann WP Fendler BJ Krause German Multicenter Study Investigating 177Lu-PSMA-617 Radioligand Therapy in Advanced Prostate Cancer Patients J Nucl Med 2017 58 1 85 90 DOI:10.2.967/jnumed.116.1.83194
[223]
A Brauer LS Grubert W Roll AJ Schrader M Schafers M Bogemann K Rahbar 177Lu-PSMA-617 radioligand therapy and outcome in patients with metastasized castration-resistant prostate cancer Eur J Nucl Med Mol Imaging 2017 44 10 1663 1670 DOI:10.1.007/s00259-017-3751-z
[224]
MS Hofman J Violet RJ Hicks J Ferdinandus SP Thang T Akhurst A Iravani G Kong A Ravi Kumar DG Murphy P Eu P Jackson M Scalzo SG Williams S Sandhu ((177)Lu)-PSMA-617 radionuclide treatment in patients with metastatic castration-resistant prostate cancer (LuPSMA trial): a single-centre, single-arm, phase 2 study Lancet Oncol 2018 19 6 825 833 DOI:10.1.016/S1470-2045(18)30198-0
[225]
Lutetium-177 (Lu177) Prostate-Specific Antigen (PSMA)-Directed EndoRadiotherapy In, ClinicalTrials.gov, Identifier: NCT03042312 2017
[226]
Study of 177Lu-PSMA-617 In Metastatic Castrate-Resistant Prostate Cancer (VISION) In, ClinicalTrials.gov, Identifier: NCT03511664 2018
[227]
MD Henry S Wen MD Silva S Chandra M Milton PJ Worland A prostate-specific membrane antigen-targeted monoclonal antibody-chemotherapeutic conjugate designed for the treatment of prostate cancer Cancer Res 2004 64 21 7995 8001 DOI:10.1.158/0008-5472.CAN-04-1722
[228]
MI Milowsky MD Galsky MJ Morris DJ Crona DJ George R Dreicer K Tse J Petruck IJ Webb NH Bander DM Nanus HI Scher Phase 1/2 multiple ascending dose trial of the prostate-specific membrane antigen-targeted antibody drug conjugate MLN2704 in metastatic castration-resistant prostate cancer Urol Oncol 2016 34 12 530.e15 530.e21 DOI:10.1.016/j.urolonc.2016.0.7.0.05
[229]
D Ma CE Hopf AD Malewicz GP Donovan PD Senter WF Goeckeler PJ Maddon WC Olson Potent antitumor activity of an auristatin-conjugated, fully human monoclonal antibody to prostate-specific membrane antigen Clin Cancer Res 2006 12 8 2591 6 DOI:10.1.158/1078-0432.CCR-05-2107
[230]
VA DiPippo WC Olson HM Nguyen LG Brown RL Vessella E Corey Efficacy studies of an antibody-drug conjugate PSMA-ADC in patient-derived prostate cancer xenografts Prostate 2015 75 3 303 13 DOI:10.1.002/pros.22916
[231]
An Open-label Extension Study of PSMA ADC 2301 in mCRPC In, ClinicalTrials.gov, Identifier: NCT02020135 2017
[232]
G Fracasso G Bellisola S Cingarlini D Castelletti T Prayer-Galetti F Pagano G Tridente M Colombatti Anti-tumor effects of toxins targeted to the prostate specific membrane antigen Prostate 2002 53 1 9 23 DOI:10.1.002/pros.10117
[233]
P Wolf D Gierschner P Buhler U Wetterauer U Elsasser-Beile A recombinant PSMA-specific single-chain immunotoxin has potent and selective toxicity against prostate cancer cells Cancer Immunol Immunother 2006 55 11 1367 73 DOI:10.1.007/s00262-006-0131-0
[234]
P Wolf K Alt D Wetterauer P Buhler D Gierschner A Katzenwadel U Wetterauer U Elsasser-Beile Preclinical evaluation of a recombinant anti-prostate specific membrane antigen single-chain immunotoxin against prostate cancer J Immunother 2010 33 3 262 71 DOI:10.1.097/CJI.0b013e3181c5495c
[235]
N Krall J Scheuermann D Neri Small targeted cytotoxics: current state and promises from DNA-encoded chemical libraries Angew Chem Int Ed Engl 2013 52 5 1384 402 DOI:10.1.002/anie.201204631
[236]
Phase 1 of EC1169 In Patients With Recurrent MCRPC In, ClinicalTrials.gov, Identifier: NCT02202447 2014
[237]
M Wu Y Wang Y Wang M Zhang Y Luo J Tang Z Wang D Wang L Hao Z Wang Paclitaxel-loaded and A10-3.2. aptamer-targeted poly(lactide-co-glycolic acid) nanobubbles for ultrasound imaging and therapy of prostate cancer Int J Nanomedicine 2017 12 5313 5330 DOI:10.2.147/IJN.S136032
[238]
M Pan W Li J Yang Z Li J Zhao Y Xiao Y Xing X Zhang W Ju Plumbagin-loaded aptamer-targeted poly D,L-lactic-co-glycolic acid-b-polyethylene glycol nanoparticles for prostate cancer therapy Medicine (Baltimore) 2017 96 30 e7405 DOI:10.1.097/MD.0000000000007405
[239]
Z Chen Z Tai F Gu C Hu Q Zhu S Gao Aptamer-mediated delivery of docetaxel to prostate cancer through polymeric nanoparticles for enhancement of antitumor efficacy Eur J Pharm Biopharm 2016 107 130 41 DOI:10.1.016/j.ejpb.2016.0.7.0.07
[240]
J Jiao Q Zou MH Zou RM Guo S Zhu Y Zhang Aptamer-modified PLGA nanoparticle delivery of triplex forming oligonucleotide for targeted prostate cancer therapy Neoplasma 2016 63 4 569 75 DOI:10.4.149/neo_2016_410
[241]
JC Leach A Wang K Ye S Jin A RNA-DNA Hybrid Aptamer for Nanoparticle-Based Prostate Tumor Targeted Drug Delivery Int J Mol Sci 2016 17 3 380 DOI:10.3.390/ijms17030380
[242]
JP Dassie LI Hernandez GS Thomas ME Long WM Rockey CA Howell Y Chen FJ Hernandez XY Liu ME Wilson LA Allen DA Vaena DK Meyerholz PH Giangrande Targeted inhibition of prostate cancer metastases with an RNA aptamer to prostate-specific membrane antigen Mol Ther 2014 22 11 1910 22 DOI:10.1.038/mt.2014.1.17
[243]
SM Taghdisi NM Danesh A Sarreshtehdar Emrani K Tabrizian M Zandkarimi M Ramezani K Abnous Targeted delivery of Epirubicin to cancer cells by PEGylated A10 aptamer J Drug Target 2013 21 8 739 44 DOI:10.3.109/1061186X.2013.8.12095
[244]
Y Gao C Zhang Y Zhou J Li L Zhao Y Li Y Liu X Li Endosomal pH-Responsive Polymer-Based Dual-Ligand-Modified Micellar Nanoparticles for Tumor Targeted Delivery and Facilitated Intracellular Release of Paclitaxel Pharm Res 2015 32 8 2649 62 DOI:10.1.007/s11095-015-1650-1
[245]
AK Pearce JD Simpson NL Fletcher ZH Houston AV Fuchs PJ Russell AK Whittaker KJ Thurecht Localised delivery of doxorubicin to prostate cancer cells through a PSMA-targeted hyperbranched polymer theranostic Biomaterials 2017 141 330 339 DOI:10.1.016/j.biomaterials.2017.0.7.0.04
[246]
F Karandish MK Haldar S You AE Brooks BD Brooks B Guo Y Choi S Mallik Prostate-Specific Membrane Antigen Targeted Polymersomes for Delivering Mocetinostat and Docetaxel to Prostate Cancer Cell Spheroids ACS Omega 2016 1 5 952 962 DOI:10.1.021/acsomega.6b00126
[247]
J Jin B Sui J Gou J Liu X Tang H Xu Y Zhang X Jin PSMA ligand conjugated PCL-PEG polymeric micelles targeted to prostate cancer cells PLoS One 2014 9 11 e112200 DOI:10.1.371/journal.pone.0112200
[248]
J Liu P Kopeckova H Pan M Sima P Buhler P Wolf U Elsasser-Beile J Kopecek Prostate-cancer-targeted N-(2-hydroxypropyl)methacrylamide copolymer/docetaxel conjugates Macromol Biosci 2012 12 3 412 22 DOI:10.1.002/mabi.201100340
[249]
ZH Peng M Sima ME Salama P Kopeckova J Kopecek Spacer length impacts the efficacy of targeted docetaxel conjugates in prostate-specific membrane antigen expressing prostate cancer J Drug Target 2013 21 10 968 80 DOI:10.3.109/1061186X.2013.8.33207
[250]
A Bandekar C Zhu R Jindal F Bruchertseifer A Morgenstern S Sofou Anti-prostate-specific membrane antigen liposomes loaded with 225Ac for potential targeted antivascular alpha-particle therapy of cancer J Nucl Med 2014 55 1 107 14 DOI:10.2.967/jnumed.113.1.25476
[251]
JB Lee K Zhang YY Tam J Quick YK Tam PJ Lin S Chen Y Liu JK Nair I Zlatev KG Rajeev M Manoharan PS Rennie PR Cullis A Glu-urea-Lys Ligand-conjugated Lipid Nanoparticle/siRNA System Inhibits Androgen Receptor Expression In vivo Mol Ther Nucleic Acids 2016 5 e348 DOI:10.1.038/mtna.2016.4.3
[252]
Y Patil H Shmeeda Y Amitay P Ohana S Kumar A Gabizon Targeting of folate-conjugated liposomes with co-entrapped drugs to prostate cancer cells via prostate-specific membrane antigen (PSMA) Nanomedicine 2018 14 4 1407 1416 DOI:10.1.016/j.nano.2018.0.4.0.11
[253]
Z Chen MF Penet S Nimmagadda C Li SR Banerjee PT Winnard Jr D Artemov K Glunde MG Pomper ZM Bhujwalla PSMA-targeted theranostic nanoplex for prostate cancer therapy ACS Nano 2012 6 9 7752 7762 DOI:10.1.021/nn301725w
[254]
SR Banerjee CA Foss A Horhota M Pullambhatla K McDonnell S Zale MG Pomper 111In- and IRDye800CW-Labeled PLA-PEG Nanoparticle for Imaging Prostate-Specific Membrane Antigen-Expressing Tissues Biomacromolecules 2017 18 1 201 209 DOI:10.1.021/acs.biomac.6b01485
[255]
Y Chen S Chatterjee A Lisok I Minn M Pullambhatla B Wharram Y Wang J Jin ZM Bhujwalla S Nimmagadda RC Mease MG Pomper A PSMA-targeted theranostic agent for photodynamic therapy J Photochem Photobiol B 2017 167 111 116 DOI:10.1.016/j.jphotobiol.2016.1.2.0.18
[256]
P Sacha T Knedlik J Schimer J Tykvart J Parolek V Navratil P Dvorakova F Sedlak K Ulbrich J Strohalm P Majer V Subr J Konvalinka iBodies: Modular Synthetic Antibody Mimetics Based on Hydrophilic Polymers Decorated with Functional Moieties Angew Chem Int Ed Engl 2016 55 7 2356 60 DOI:10.1.002/anie.201508642
[257]
J Neburkova F Sedlak J Zackova Suchanova L Kostka P Sacha V Subr T Etrych P Simon J Barinkova R Krystufek H Spanielova J Forstova J Konvalinka P Cigler Inhibitor-GCPII Interaction: Selective and Robust System for Targeting Cancer Cells with Structurally Diverse Nanoparticles Mol Pharm 2018 DOI:10.1.021/acs.molpharmaceut.7b00889
[258]
FD Meng S Wang YH Jiang CG Sui Antitumor effect of dendritic cells transfected with prostate-specific membrane antigen recombinant adenovirus on prostate cancer: An in vitro study Mol Med Rep 2016 13 3 2124 34 DOI:10.3.892/mmr.2016.4.754
[259]
HB Xi GX Wang B Fu WP Liu Y Li Survivin and PSMA Loaded Dendritic Cell Vaccine for the Treatment of Prostate Cancer Biol Pharm Bull 2015 38 6 827 35 DOI:10.1.248/bpb.b14-00518
[260]
V Baum P Buhler D Gierschner D Herchenbach GJ Fiala WW Schamel P Wolf U Elsasser-Beile Antitumor activities of PSMAxCD3 diabodies by redirected T-cell lysis of prostate cancer cells Immunotherapy 2013 5 1 27 38 DOI:10.2.217/imt.12.1.36
[261]
JT Patterson J Isaacson L Kerwin G Atassi R Duggal D Bresson T Zhu H Zhou Y Fu GF Kaufmann PSMA-targeted bispecific Fab conjugates that engage T cells Bioorg Med Chem Lett 2017 27 24 5490 5495 DOI:10.1.016/j.bmcl.2017.0.9.0.65
[262]
K Youlin Z Li G Xin X Mingchao L Xiuheng W Xiaodong Enhanced function of cytotoxic T lymphocytes induced by dendritic cells modified with truncated PSMA and 4-1BBL Hum Vaccin Immunother 2013 9 4 766 72 DOI:10.4.161/hv.23116
[263]
SP Santoro S Kim GT Motz D Alatzoglou C Li M Irving DJ Powell Jr G Coukos T cells bearing a chimeric antigen receptor against prostate-specific membrane antigen mediate vascular disruption and result in tumor regression Cancer Immunol Res 2015 3 1 68 84 DOI:10.1.158/2326-6066.CIR-14-0192
[264]
I Serganova E Moroz I Cohen M Moroz M Mane J Zurita L Shenker V Ponomarev R Blasberg Enhancement of PSMA-Directed CAR Adoptive Immunotherapy by PD-1/PD-L1 Blockade Mol Ther Oncolytics 2017 4 41 54 DOI:10.1.016/j.omto.2016.1.1.0.05
[265]
RP Junghans Q Ma R Rathore EM Gomes AJ Bais AS Lo M Abedi RA Davies HJ Cabral AS Al-Homsi SI Cohen Phase I Trial of Anti-PSMA Designer CAR-T Cells in Prostate Cancer: Possible Role for Interacting Interleukin 2-T Cell Pharmacodynamics as a Determinant of Clinical Response Prostate 2016 76 14 1257 70 DOI:10.1.002/pros.23214
[266]
Q Zhang BT Helfand BA Carneiro W Qin XJ Yang C Lee W Zhang FJ Giles M Cristofanilli TM Kuzel Efficacy Against Human Prostate Cancer by Prostate-specific Membrane Antigen-specific, Transforming Growth Factor-beta Insensitive Genetically Targeted CD8(+) T-cells Derived from Patients with Metastatic Castrate-resistant Disease Eur Urol 2018 73 5 648 652 DOI:10.1.016/j.eururo.2017.1.2.0.08
[267]
Y Langut A Talhami S Mamidi A Shir M Zigler S Joubran A Sagalov E Flashner-Abramson N Edinger S Klein A Levitzki PSMA-targeted polyinosine/polycytosine vector induces prostate tumor regression and invokes an antitumor immune response in mice Proc Natl Acad Sci U S A 2017 114 52 13655 13660 DOI:10.1.073/pnas.1714587115
[268]
K Muthumani L Marnin SB Kudchodkar A Perales-Puchalt H Choi S Agarwal VL Scott EL Reuschel FI Zaidi EK Duperret MC Wise KA Kraynyak KE Ugen NY Sardesai J Joseph Kim DB Weiner Novel prostate cancer immunotherapy with a DNA-encoded anti-prostate-specific membrane antigen monoclonal antibody Cancer Immunol Immunother 2017 DOI:10.1.007/s00262-017-2042-7
Share
Back to top