Next Article in Journal
Non-Invasive Profiling of Advanced Prostate Cancer via Multi-Parametric Liquid Biopsy and Radiomic Analysis
Next Article in Special Issue
A Bright Horizon: Immunotherapy for Pediatric T-Cell Malignancies
Previous Article in Journal
Protein Expression of AEBP1, MCM4, and FABP4 Differentiate Osteogenic, Adipogenic, and Mesenchymal Stromal Stem Cells
Previous Article in Special Issue
Rho-Kinase as a Target for Cancer Therapy and Its Immunotherapeutic Potential
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Immunotherapy in Advanced Prostate Cancer—Light at the End of the Tunnel?

1
Department of Hematology and Oncology, University Medical Center Hamburg-Eppendorf, Martinistrasse 52, 20246 Hamburg, Germany
2
Martini-Klinik, Prostate Cancer Center, University Medical Center Hamburg-Eppendorf, Martinistrasse 52, 20246 Hamburg, Germany
3
Department of Tumor Biology, University Medical Center Hamburg-Eppendorf, Martinistrasse 52, 20246 Hamburg, Germany
4
Laboratory of Pharmacology, A.V. Zhirmunsky National Scientific Center of Marine Biology, Palchevskogo Str. 17, 690041 Vladivostok, Russia
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2022, 23(5), 2569; https://doi.org/10.3390/ijms23052569
Submission received: 3 February 2022 / Revised: 22 February 2022 / Accepted: 24 February 2022 / Published: 25 February 2022
(This article belongs to the Special Issue Immunotherapies for Cancer)

Abstract

:
Immunotherapeutic treatment approaches are now an integral part of the treatment of many solid tumors. However, attempts to integrate immunotherapy into the treatment of prostate cancer have been disappointing so far. This is due to a highly immunosuppressive, “cold” tumor microenvironment, which is characterized, for example, by the absence of cytotoxic T cells, an increased number of myeloid-derived suppressor cells or regulatory T cells, a decreased number of tumor antigens, or a defect in antigen presentation. The consequence is a reduced efficacy of many established immunotherapeutic treatments such as checkpoint inhibitors. However, a growing understanding of the underlying mechanisms of tumor–immune system interactions raises hopes that immunotherapeutic strategies can be optimized in the future. The aim of this review is to provide an overview of the current status and future directions of immunotherapy development in prostate cancer. Background information on immune response and tumor microenvironment will help to better understand current therapeutic strategies under preclinical and clinical development.

Graphical Abstract

1. Background

Prostate cancer (PCa) is the most common malignancy and second leading cause of cancer-related death amongst men in the Western world [1].
In the metastatic setting, tremendous progress has been made in recent years. Thus, in metastatic hormone-naïve PCa (mHNPC), combinational treatment with androgen-deprivation therapy (ADT) and new hormonal agents (NHA) or chemotherapy with docetaxel is recommended [2]. Interesting new data on triple therapy consisting of ADT, docetaxel, and abiraterone or darolutamide compared with hormone chemotherapy alone showed a clinically significant survival benefit and will set a new standard, especially for patients with high-risk constellations [3,4]. In the castration-resistant setting, additional treatment options include cabazitaxel, PARP inhibitors, and PSMA radioligand therapy [5,6,7,8]. However, prognosis of advanced-stage patients is still poor due to the development of resistance to currently available standard therapies. Therefore, new therapeutic approaches are urgently needed.
The idea to utilize the patients’ immune system to fight tumors has revolutionized the field of anticancer therapy within the last decade [9]. In fact, immunotherapeutic approaches have been triumphant in several highly immunogenic cancers, often called “hot tumors” (such as melanoma, renal carcinoma, and lung cancer, among others) [10,11,12]. Thus, immune-checkpoint monotherapies or combination regimens targeting cytotoxic T lymphocyte antigen 4 (CTLA-4) and/or the programmed death-1 (PD-1)/PD-1 ligand (PD-L1) axis have become an integral part of various first-line standard therapies in a variety of malignancies. In contrast, so-called “cold tumors”, such as prostate cancer (PCa), possess an immunosuppressive tumor microenvironment (TME) resulting in a very restricted response to immunotherapy. In fact, immunotherapy has so far been generally disappointing in PCa. To date, despite intensive efforts, sipuleucel T is the only immunotherapeutic agent that has achieved a significant survival benefit in a randomized Phase 3 clinical trial (see Section 3.1.1). [13]. Currently investigations are conducted to make the immunologically cold PCa accessible to immunotherapy by manipulating the tumor microenvironment as well as implementing new immunological treatment strategies with bispecific T cell engagers (BiTEs) or CAR-T cells.
In this review, we discuss mechanisms contributing to immune response and resistance of PCa, summarize the different treatment approaches and study results available, and provide an overview of the current study landscape.

2. Immune Response and the Role of the Microenvironment in Prostate Cancer

2.1. Intrinsic Factors Influencing Immune Response

2.1.1. Tumor Mutational Burden and Neoantigen Expression

The recognition of neoantigens is a central mechanism mediating antitumor immunity (Figure 1). Neoantigens result from non-synonymous mutations translating into aberrant proteins. These are presented to the immune system, which consequently recognizes the tumor cell as “foreign”. Multiple studies have confirmed that response to immunotherapy is associated with tumor mutational burden (TMB) in a variety of tumor entities [14]. TMB in PCa is low, especially in comparison to immunological “hot” tumors such as melanoma or non-small cell lung cancer [15]. The resulting low number of neoantigens is considered one of the tumor intrinsic factors contributing to the low response rates to immunotherapy in PCa [15]. Although TMB has been shown to be a good predictor to immunotherapy response, there are still patients with antitumor responses despite a divergent TMB result (Figure 1). In addition, PCa is characterized by a comparatively high amount of structural variants, such as indels and insertions or fusions, which also lead to expression of neoantigens [16]. A recent study demonstrated a correlation of high-fusion burden in PCa with increased immune infiltration, PD-L1 expression on immune cells, and immune signatures, representing activation of T cells and M1 macrophages [17].

2.1.2. Expression of Programmed Death Ligand-1 (PD-L1)

Generally, PD-L1 expression levels in PCa are lower compared with other cancers, although up to one-third of mCRPC tumors may show some PD-L1 expression on tumor cells [18]. The range of reported PD-L1 expression in PCa is wide, varying from no expression to over 90% in some patients. For example, immunohistochemical expression of PD-L1 has been detected in 29% acinar PCas, 7% ductal PCas, and 46% neuroendocrine PCas [19]. PD-L1 expression on TILs was found in 9%–14.6% of the cases [18,20,21,22,23,24,25]. While some studies observed a correlation between Gleason Score and PD-L1 expression [21,23], this was not confirmed by others [18,22,24]. Data on the correlation of PD-L1 expression with gene alterations related to tumor progression and aggressiveness are not yet comprehensively available (reviewed by Palicelli et al. [26]). Indeed, in patients with MSI-H/dMMR-positive disease, PD-L1 expression was detected in only about 12%. Patients with a PTEN deletion carry increased PD-L1 expression in about 10%, whereas some enrichment is apparently found in SPOP-mutated PCa. Data on expression in homologous recombination repair (HRR) defects remain to be comprehensively collected [26]. Although study results have been inconsistent, increased PD-L1 expression has been associated with a higher risk of biochemical recurrence or metastatic progression [21,24,27,28]. This substantial variability can also be found in PD-L1 as a predictor of immunotherapeutic responses (Figure 1).

2.1.3. DNA Repair Defects

Loss of function of DNA damage repair (DDR) genes potentiates TMB and genomic instability. The mismatch repair (MMR) system repairs single base substitutions and short indels, for example during DNA synthesis. Defects in the MMR system can lead to point mutations, frameshifts, and the phenomenon of microsatellite instability (MSI). A recent study found 3% of MSI high tumors in an unselected cohort of PCa patients. This percentage can rise up to 12% in advanced PCa [29,30]. Indeed, MSI seems to be acquired in late disease stages, while about 22% of affected patients already harbor germline mutations [29]. MMR-deficient PCa shows higher immune infiltration and an increased response to immunotherapy compared to MMR-proficient tumors [29,31,32,33] (Figure 1). Still, about half of the patients with high MSI do not respond to immune therapy. However, causes of this primary resistance remain to be examined [29].
Mutations in other DDR genes occur more frequently. These include alterations in the HRR pathway genes, e.g., BRCA2 and ATM, among others [34]. Analysis of patient tissue revealed that BRCA2-mutated tumors had more T cells within the tumor compared to the extratumoral tissue. However, these tumors were simultaneously infiltrated with more regulatory T cells (Treg) compared to BRCA2 wild-type tumors [35]. Mechanistically, impaired DNA repair due to BRCA2 loss leads to cytosolic DNA fragments. In turn, these activate a cGAS/STING-mediated interferon response [36]. However, the inconsistency surrounding clinical outcomes in patients bearing cancers with DDR alterations and treated with immune checkpoint inhibitors questions its relevance as a biomarker [37,38].

2.1.4. Inactivation of PTEN

The loss of tumor-suppressor protein PTEN, especially in combination with inactivation of RB1 and/or TP53, has been recognized as one of the signs of aggressive variant PCa (AVPC) [39,40]. PTEN inactivation is present in approximately 20% of primary tumors and in around 40% of advanced PCa [41]. In immune response, functional PTEN has been associated with activation of pro-inflammatory INF1 and NF-kB pathways [42,43]. In line with this, PTEN-deficient PCa tumors have higher Treg cell infiltration, which is associated with an immunosuppressive TME of these tumors as well as resistance to immunotherapy [44] (Figure 1). Recently, increased recruitment of myeloid-derived suppressor cells (MDSCs) has been described in PTEN-deficient tumors [45,46]. Based on these findings, targeting of MDSCs in PCa tumors expressing mutant or null PTEN is currently under clinical investigations.

2.1.5. Androgen Receptor Signaling

Androgen receptor (AR) signaling is critical for progression and survival of normal and malignant prostate cells [47]. ADT decreases the levels of circulating testosterone (via e.g., inhibition of the GnRH receptor) and is a central component of systemic therapy for advanced PCa. Direct inhibition of the AR or of androgen synthesis are additional therapeutic approaches to successfully influence AR signal transduction [48].
The effect of ADT on tumor immunogenicity is complex and, in some cases, controversial. In AR-dependent tumors, an androgen withdrawal initially results in an increase of TILs and reduction of Treg cells [49,50,51]. In vivo studies have indicated direct immunomodulatory effects of ADT on PCa TME, though the results were strongly dependent on the utilized models [49,50]. Moreover, development of castration resistance under continuous ADT treatment seems to correlate with simultaneous increase of immune tolerance [52]. Additionally, ADT can impair adaptive immune response by inhibition of T cell function, which may be related to the off-target effects of AR antagonists on the γ-aminobutyric acid receptor [53]. Thus, in patients suffering from localized PCa, treatment with ADT in combination with GVAX (cell-based vaccine) resulted in promoted infiltration of CD8 + T cells accompanied by a simultaneous increase of the immunosuppressive Treg cell population [54]. In addition, the upregulation of immune checkpoints and TILs has been observed in samples obtained from patients receiving ADT therapy [55].

2.2. The Role of Tumor Microenvironment in Prostate Cancer

The TME of PCa is highly immunosuppressive due to infiltration of regulatory T cells, tumor-associated macrophages (TAM), and MDSCs, the cytokine milieu secreted by tumor stromal cells and fibroblasts, as well as the production of adenosine via prostatic acid phosphatase [56,57].

2.2.1. The Tumor Cytokine Milieu

Cytokines regulate and facilitate immune response to different stimuli, including tumor development [58]. In PCa progression as well as during PCa, directed immunotherapy cytokines play a dual role [59]. On the one hand, in local chronic inflammatory conditions the production of pro-inflammatory cytokines leads to the promotion of CTL infiltration into the tumor; on the other hand, the traffic of immunosuppressive MDSC and Treg cells is increased due to the stimulation by IL-1β and IL-2 cytokines (reviewed in [59]). Moreover, tumor necrosis factor-α (TNF-α) and IL-17 were found to induce expression of immunosuppressive PD-L1 on PCa cell surface contributing to immune resistance [60]. Accordingly, increased expression of anti-inflammatory cytokines in PCa was accompanied by substantial suppression of CTL infiltration and activity [61]. Thus, the levels of anti-inflammatory cytokines IL-4, IL-6, and IL-10 were elevated in serum of patients suffering from hormone-refractory PCa [62] and associated with increased PSA levels. Additionally, it has been shown in patients that secretion of immunosuppressive cytokines such as TGF-β and IL-10 facilitate a weak antitumor immune response and condition a poor treatment prognosis [61,63,64].

2.2.2. Myeloid-Derived Suppressor Cells (MDSCs)

The release of inflammatory cytokines leads to accumulation (recruitment) of MDSC in TME as well as differentiation of myeloid cells to MDSC [65]. MDSC are a population of immature myeloid cells that exhibit immune-suppressive effects on T cells and NK cells and therefore are considered to be an important mechanism of immunotherapeutic resistance in PCa (reviewed in [66,67]) (Figure 1). MDSCs promote recruitment of Tregs that suppress the function of CTLs, and can differentiate to monocytes or neutrophils and then further to M2 macrophages, which also inhibits CTL activity [68]. Consequently, low T cell recruitment and reduced activity result in inefficacy of immunotherapeutic strategies in PCa. Additionally, MDSC promote growth and survival of PCa cells via other mechanisms unrelated to immunosuppression [45,69]. Increased MDSC infiltration in tumor tissue and blood has been reported in patients with an accumulation from localized to metastatic PCa [70]. This correlated positively with Treg levels and negative prognostic markers [71]. Moreover, a large population of MDSCs has been found in the metastatic regions of the bones, which are the main metastatic sites of PCa [72,73]. Remarkably, the elimination of MDSC significantly improved T cells infiltration and promoted anticancer effects of immune checkpoint inhibitors (CI) in the pre-clinical setting [74]. In line with this, S100A9 inhibitor tasquinimod, a MDSC-targeting drug, exhibited a clear benefit for progression-free survival (PFS), even though it failed to improve overall survival (OS) in mCRPC [75].

2.2.3. Tumor-Associated Macrophages (TAMs)

TAMs are a key component of the inflammatory TME. These highly heterogeneous cells originate either from resident tissue-specific macrophages or newly recruited monocytes [76]. Generally, they can be classified in a tumor-inhibiting M1 phenotype or a tumor-promoting M2 phenotype. However, mixed phenotypes as well as a phenotype switching upon stimulation including the development of subpopulations with differences in antigen expression have been observed [77]. TAMs can stimulate tumor cell proliferation, migration and genetic instability [78]. In PCa, an increased TAM density has been associated with higher Gleason score and shorter cancer-specific survival [77,79,80]. Of note, increased activation of osteoclast related pathways has been associated with TAMs as well [81]. This is of great interest because PCa metastasizes mainly to the bones [82]. A targeted approach of TAMs seems to be a reasonable therapeutic procedure and is currently being investigated mainly in pre-clinical models [83].

2.2.4. Stromal Cells

TME consists of various populations of stromal cells. Apart from innate and adaptive immune cells, it includes non-immune cell like fibroblasts, endothelial cells, adipocytes, mesenchymal stromal cells (MSCs), and pericytes [84]. In terms of tumorigenesis, the cancer-associated fibroblasts (CAFs) and endothelial cells have been assumed to be most relevant, with the latter being involved in promotion of neo-angiogenesis, and thus local progression and hematogenous dissemination [85]. CAFs are the most abundant stromal cells represented in the TME [86]. These cells may be of a different origin and are known to play an important role in tumorigenesis via promotion of angiogenesis, remodeling of the extracellular matrix (which supports tumor cells invasion), and expression of growth factors as well as cyto- and chemokine, thereby mediating a cross-talk with PCa cells [86,87]. The most important factors involved in this interaction are bFGF, PDGF, and TNF-α, as well as MMPs and VEGF. These mediators promote acquisition of stem cell properties as well as an epithelial–mesenchymal transition (EMT) of the PCa cells, ultimately leading to a more aggressive phenotype and increased drug resistance [88]. Moreover, the proliferation of CAFs leads to an alteration of vascular structure of the tumor and the surrounding tissues, which in turn results in development of hypoxia and local inflammation, and consequently leads to the infiltration of immunosuppressive MDSCs and Treg cells in TME [89].
Additionally, CAFs can directly inhibit the antitumor functions of cytotoxic lymphocytes. Thus, CAFs abrogate the cytotoxic function of NK cells due to the expression of PGE2 and IDO by CAFs, which leads to the inhibition of cytotoxic function and inactivation of NK cells [90,91]. Additionally, CAFs express MMPs, which degrade MICA/B on the surface cancer cell (the NKG2D ligands) and therefore abrogate a NKG2-dependent cytotoxicity of NK cells [90]. CAFs can also inhibit the function of cytotoxic T cells via various mechanisms, such as the following: (a) IL-1-mediated expression of PD-L1; (b) overexpression of CD39 and CD73, which results in generation of high amounts of immunosuppressive adenine in the TME; (c) release of lactate by the glycolytic CAFs, which affects polarization and function of cytotoxic T lymphocytes; and (d) activation of TGF-β signaling, which is correlated with the exhaustion of T cells and was associated with poor response to anti-PD-1 therapy [87] (Figure 1).

2.2.5. Adenosine in PCa

Adenosine is a small molecule, which, among others, acts as an anti-inflammatory mediator. This effect is executed mainly via adenosine-dependent activation of A2a/A2b receptors expressed on the surface of T cells resulting in their inactivation. In normal settings, this mechanism helps to protect healthy tissues from the damage induced by the inflammatory processes [92]; however, within the tumor microenvironment, the immunosuppressive effects of adenosine results in tumor immune evasion [93]. Thus, adenosine suppresses the cytotoxic function of TILs [93,94]; additionally, it can increase Treg cells and MDSCs, increase the number of tumor-promoting and angiogenic fibroblasts, and inhibit functionality of dendritic cells [95,96,97,98].
In physiological conditions, adenosine is produced from ATP via its consecutive dephosphorylation catalyzed by ectonucleotidases like CD39, CD73, prostatic acid phosphatase (PAP), and alkaline phosphatase [99]. Interestingly, extracellular ATP functions such as the danger-associated molecular pattern (DAMP) promotes immune responses, while in its dephosphorylated form adenosine exhibits immunosuppressive properties, therefore stabilizing and balancing an immune response [98]. Of note, PAP is highly expressed in PCa tissues [100]. Therefore, PAP targeting may represent an attractive mechanism to increase the efficacy of immunotherapy by the inhibition of adenosine production. However, vaccination trials targeting PAP have been rather disappointing (see below). Early clinical trials are now investigating efficacy of the A2a and/or A2b receptor antagonists and CD73 ectonucleotidase inhibitors in PCa [101,102].

3. Immunotherapeutic Treatment Approaches

3.1. Vaccine-Based Treatment Modalities

Vaccination strategies are based on activation of the immune system against specific tumor-associated antigens (TAA), e.g., prostate-specific antigen (PSA), prostate-specific membrane antigen (PSMA), prostatic acid phosphatase (PAP), or prostate stem cell antigen (PSCA) [103,104]. Applied with co-stimulatory molecules or loaded on patients’ immune cells, vaccines are recognized by APC cells, processed, and presented with MHC class 1 molecules to CD8 + cells, turning them into cytotoxic T effector cells (Figure 2) [105]. Various vaccination approaches have been tested in PCa to date including cell-based vaccines, peptide vaccines, viral/bacterial vaccines, as well as DNA- or RNA-based vaccines.

3.1.1. Cell-Based Vaccines

Cell-based vaccines are derived from patients’ own tumor cells (autologous) or from tumor cell lines (allogeneic). For example, the granulocyte-macrophage colony-stimulating factor-transduced allogeneic prostate cancer cell vaccine (GVAX) was synthesized from two prostate cancer cell lines, LNCaP and PC-3, and transfected with a human GM-CSF gene. However, GVAX failed to succeed in two phase 3 clinical trials applied as monotherapy or in combination with docetaxel (VITAL-1; VITAL-2) [106]. In contrast, sipuleucel-T, a dendritic cell (DC)-based immunotherapy, improved overall survival (OS) in PCa and was approved by the FDA in 2010 for the treatment of asymptomatic or minimally symptomatic mCRPC. For the synthesis of sipuleucel-T, patient-derived peripheral blood mononuclear cells (PBMCs) are stimulated ex vivo with a recombinant fusion protein consisting of prostatic acid phosphatase and the granulocyte-macrophage colony-stimulating factor. Afterwards, the PBMCs are reinfused every 2 weeks for a total of three infusions. Results from the phase 3 study IMPACT showed a relative reduction in the risk of death of 22% (HR 0.78; 95% CI 0.61 to 0.98; p = 0.03) and a significant improvement of 4.1 months in median OS in the sipuleucel-T group compared to the placebo group. Immune responses to the immunizing antigen were detected after sipuleucel-T treatment [13]. However, in Europe, the marketing authorization for sipuleucel-T was withdrawn at the request of the marketing authorization holder, limiting its clinical application. Currently, different combinational approaches are under investigation, including promising results reported on a combination with radium-223 (NCT02463799) [107]. Despite this initial success, further DC vaccines have failed to succeed [108]. For example, in the phase 3 VIABLE trial, the addition of autologous DC-based immunotherapy to docetaxel and prednisone did not prolong overall survival compared to chemotherapy alone in mCRPC patients [109]. In addition, no differences in the secondary efficacy end points (rPFS, time to PSA progression, or skeletal-related events) were reported.

3.1.2. Peptide Vaccines

Peptide vaccines are produced by using specific epitope subunits of TAA. Comparable to cell-based vaccinations, peptide vaccination strategies can be based on common TAA or individual peptide structures. To date, different peptide vaccines have been evaluated in early phase 1/2 clinical trials, with some thoroughly promising initial results. For example, targeting Ras homolog gene family member C (RhoC) induced a potent and long-lasting T cell immunity in the majority of patients who had previously undergone radical prostatectomy (NCT03199872) [110]. Additionally, human telomerase reverse transcriptase (hTERT) peptide vaccine UV1 in combination with GM-CSF induced specific immune responses in the majority of mHNPC patients unselected for HLA type with tolerable adverse events [111]. Additional early phase clinical trials are currently recruiting, e.g., with B-cell lymphoma extra-large protein (Bcl-xl) 42-CAF09b (Bcl-xl_42: peptide fragment; CAF09b: adjuvance to enhance immunostimualtion) (NCT03412786; Table 1).
Personalized peptide vaccination (PPV) uses multiple cancer peptides based on the pre-existing host immunity. In a randomized phase 2 trial, HLA-type specific peptides were chosen for vaccination based on the evaluation of both antipeptide IgG levels in plasma and CTL precursors in PBMCs of each patient (a maximum of four reactive proteins) and compared to low dose dexamethasone. Remarkably, PSA-PFS and median OS were significantly longer in the vaccination group compared to the dexamethasone group (PSA-PFS: 22.0 vs. 7.0 months; p = 0.0076; OS: 73.9 vs. 34.9 months; p = 0.00084), respectively [112]. Currently, a phase 1 study is evaluating PPV in combination with immune modulator Poly-ICLC, and hematopoietic cytokine CDX-301 in the adjuvant setting (Table 1; NCT05010200).

3.1.3. Viral/Bacterial-Based Vaccines

PROSTVAC (PSA-TRICOM) is a viral vector-based vaccine using two different poxviral vectors for human PSA (PROSTVAC-V and -F). Additionally, these vectors include three co-stimulatory molecules for T cells (TRICOM) to enhance immune response. However, promising results of a phase 2 study [113] did not transfer into the phase 3 setting (NCT01322490), which compared patients receiving PROSTVAC (n = 432) or PROSTVAC plus granulocyte-macrophage colony-stimulating factor (n = 432) to placebo (n = 433). Neither treatment regimen had an impact on OS or on the number of patients alive without events [114]. Accordingly, FDA or EMA approval was not granted; however, combinational therapies are currently under investigation. Thus, PROSTVAC combined with CIs ipilimumab/nivolumab as well as a the neoantigen DNA vaccine are currently evaluated in mHNPC in a phase 1 clinical trial (NCT03532217). In addition, the efficacy of PROSTVAC co-applicated with or followed by docetaxel is determined in first-line treatment of mHSPC patients in a phase 2 clinical trial (NCT02649855).
In addition to viruses, bacterial microorganisms can also serve as a source for vaccines. Therapeutic approaches using live tumor-targeting bacteria can either be applied as a monotherapy or complement other anticancer therapies to achieve better clinical outcomes [115]. In general, a major concern in the field of bacterial-based cancer therapies (BBCT) is toxicity due to associated toxins, which may cause side effects, similar to infections [116].
In the phase 1/2 trial KEYNOTE-046, DXS-PSA, an attenuated Listeria monocytogenes-based immunotherapy targeting PSA is examined in combination with pembrolizumab in mCRPC. Overall, 43% of the patients achieved a PSA reduction with a median OS of 33.6 months (NCT02325557) according to Stein et al. [117].
However, while virus-based immunotherapy is on the rise due to its relatively rapid and relatively uncomplicated production, the acceptance and implementation of BBCT is not yet at this scale. In addition to toxicity, cultural stigmas must be addressed before any decisive progress will be made. Recruiting trials on viral- and bacterial-based vaccines are shown in Table 1.

3.1.4. DNA and RNA Vaccines

DNA vaccines serve as vehicles for in vivo transfection and antigen production. They consist of a plasmid DNA that encodes the antigen of interest under the control of a eukaryotic promoter [118]. To date, most DNA vaccines are focusing on antigens specific for PCa (e.g., PAP, PSA, or AR), while targeting of other tumor-specific mutation-associated neoantigens is challenging due to the rather low TMB of PCa. A cancer vaccine containing plasmid DNA encoding human PAP (pTVG-HP) has been investigated in several phase 1/2 trials. In these trials, multiple vaccinations were required to maintain an immune response, and still most patients did not benefit [118]. Thus, in a phase 2 clinical trial on patients with non mHNPC and biochemical recurrence, only a subgroup (having a rapid PSA doubling time) was identified to have an improved outcome concerning 2-year metastases-free survival (NCT01341652) [119]. Studies using PSA or PSMA as TAAs have been similarly disappointing so far. However, a combination of INO-5150 (a synthetic DNA therapy with plasmids encoding for PSA and PSMA) and INO-9012 (a synthetic DNA plasmid encoding for IL-12) showed first promising results in a phase 1/2 trial. In fact, a correlation of CD38 + and perforin + co-positive CD8 + T cell frequency to attenuated PSA rise for patients with biochemical recurrent non mHNPC was found [120]. In addition, the DNA vaccine pTVG-AR encoding the AR ligand binding has been evaluated in a phase 1 study (NCT024117869). Patients who developed T cell immunity had a significantly prolonged PSA-PFS compared with patients without immunity (HR = 0.01; 95% CI, 0.0–0.21; p = 0.003) [121]. Based on these findings, further investigations in combination with CIs have been initiated (NCT04989946; NCT04090528; Table 1).
Rapid development potentials and cost-effective manufacturing are the advantages of mRNA vaccines [105]. However, despite promising first signals in phase 1 clinical trials, follow-up investigations have been rather disappointing. Thus, standard therapy in combination with CV9104, a sequence-optimized, free, and protamine-complexed mRNA vaccine encoding the antigens PSA, PSMA, PSCA, STEAP1, PAP, and MUC1, failed to improve OS in mildly or asymptomatic mCRPC compared to standard therapy alone [122]. Another mRNA vaccine, W_pro1, which targets five antigens expressed in de novo and metastatic PCa and stably complexed with liposomes, is currently under clinical investigation in combination with PD-1 inhibitor Cemiplimab (NCT04382898, Table 1).

3.2. Checkpoint Inhibitors

Although the initial trials of immune CI in unselected PCa patients [123,124] failed to demonstrate significant clinical benefit (Figure 2), individual PCa patients showed impressive and durable responses. This raises the hope that immunotherapy could be a potential treatment option after an appropriate biomarker-based preselection [125,126].

Checkpoint Inhibitor Monotherapy

Anti-CTLA-4 Antibodies

Treatment with the anti-CTLA-4 antibody ipilimumab failed to demonstrate an OS benefit in two phase 3 clinical trials in patients with mCRPC (CA184 043 and CA184 095). Initially, ipilimumab was evaluated in docetaxel-pretreated mCRPC patients who had at least one bone metastasis. All patients received bone-directed radiotherapy (8 Gy in one fraction) followed by four courses of ipilimumab (10mg/kg) or placebo every three weeks. The primary endpoint was not reached with a median OS of 11.2 months (95% CI 9.5–12.7) with ipilimumab versus 10.0 months (8.3–11.0) with placebo (hazard ratio (HR) 0.85, 0.72–1.00; p = 0.053). However, in long-term analyses, a piecewise hazard model showed an improvement over time with a HR of 1.46 for the first five months, but onlyn 0.6 beyond 12 months with two to three times higher survival rates at 3 years and beyond for ipilimumab [127]. Next, ipilimumab monotherapy was evaluated in mild or asymptomatic mCRPC patients without visceral metastases prior to chemotherapy. Using the same regimen as in the previous trial, there was again no survival benefit for immunotherapy with a median OS of 28.7 months (95% CI, 24.5 to 32.5 months) in the ipilimumab arm compared to 29.7 months (95% CI, 26.1 to 34.2 months) in the placebo arm (HR 1.11; 95.87% CI, 0.88 to 1.39; p = 0.3667). Nevertheless, improved PFS and PSA response rates suggested antitumor activity in a patient subset [123].
Interestingly, exceptional clinical benefit has been reported in some patients with long-term responses, raising the question for appropriate selection criteria [125]. Recently, a single-center phase 2 clinical trial was conducted to address the impact of T cell responses to cancer neoantigens for an effective antitumor response to ipilimumab in mCRPC patients. Therefore, patients were assigned to two predominant categories depending on radiographic and/or clinical PFS (rcPFS). The favorable group had a rcPFS > 6 months and an OS > 12 months (n = 9), while the unfavorable cohort was characterized by a rcPFS < 6 months and an OS < 12 months (n = 10). Remarkably, in the pretreatment tumors, the IFN-γ response pathway signature was higher in patients in the favorable cohort compared to the unfavorable cohort. In addition, the favorable cohort had higher T cell gene signatures, including those of cytotoxic and memory T cells. Of note, an increased TMB was not associated with improved clinical responses to ipilimumab in this small number of patients. T cell responses to PSMA, PAP, and cancer neoantigens were only observed in patients within the favorable cohort [128].

PD-1/PD-L1 Inhibitors

Targeting the PD-1/PD-L1 axis alone has resulted in limited success in an unselected patient population [129]. Thus, in a phase 1/2 basket trial, nivolumab showed a significantly lower response in patients with PCa (n = 17) compared with those suffering from non-small cell lung cancer, melanoma, or renal-cell carcinoma (NCT00730639) [130].
KEYNOTE-028, a phase 1b multicenter basket trial with PD-1 inhibitor pembrolizumab included patients with advanced PCa who had progressed on standard therapy and had measurable disease per RECIST v1.1 as well as a positive PD-L1 expression in ≥1% of tumor or stromal cells (n = 23). In this small cohort of patients, a remarkable ORR of 17.4% and SD of 34.8% were reported, with an average response duration of 13 months (NCT02054806) [131]. Next, pembrolizumab was evaluated in KEYNOTE-199, a five-cohort, open-label, phase 2 study. In cohorts 1–3, patients with mCRPC treated with docetaxel and one or more of the targeted endocrine therapies were enrolled (cohort 1 and 2: RECIST measurable PD-L1–positive and PD-L1–negative disease; cohort 3: bone-predominant disease, regardless of PD-L1 expression). ORR and DCR were low with 5% and 10% in cohort 1 and 3% and 9% in cohort 2, respectively. In patients with predominant bone disease a DCR of 22% was reported. Of note, comparable to other entities, responses that did occur were durable. Among the nine patients with RECIST-measurable disease who achieved complete or partial response, five had responses ongoing at data cutoff with a median duration of 16.8 months (NCT02787005) [130].
The PD-L1 antibody atezolizumab was evaluated in 35 mCRPC progressive patients on sipuleucel-T or enzalutamide in a phase 1 trial. Two patients reached a PR, with one of those patients inheriting a MSH2 and MSH6 deletion, therefore being considered as MMR deficient. A 50% PSA response was reached by 8.6%, with a general median OS of 14.7 months and a 1-year OS rate of 52.3%. Although biomarker analyses showed that atezolizumab activated immune responses, no consistent biomarker linked to treatment efficacy was found [132].
Improved efficacy of PD-1/PD-L1 inhibition has been reported in PCa patients harboring MSI high tumors, with radiographic responses in 36% and 50% PSA decline in 54% of the cases, with most patients reaching long-term disease control (6).
In contrast, data on the impact of HRD on immune response are less clear. In KEYNOTE-199, a conditionally assessable tendency of higher pembrolizumab efficacy with a 11% ORR has been reported in patients with BRCA1/2 or ATM alterations compared to 3% in men without HRD defects (3). Currently, ImmunoProst is evaluating nivolumab in HRD-positive mCRPC (NCT03040791; Table 2).
Cycline-dependent kinase 12 (CDK12) and its loss of function due to biallelic mutation (prevalent in up to 6.9% of mCRPC patients) results in severe genomic instability, followed by a high load of immunogenic neoantigens because of widespread gene fusions. In a retrospective analysis, a PSA response was reported in 33% heavily pre-treated patients treated with a PD-1 inhibitor [133]. This immunogenicity might be a potent predictive biomarker for further investigation of immunotherapeutic efficacy in mCRPC (11). The IMPACT trial is currently evaluating a combinational CTL-A4/PD-1 inhibition in metastatic cancer harboring a loss of CDK12 function (NCT03570619; Table 2).

3.3. Checkpoint Inhibitor Combinations

3.3.1. PD-1/PDL-1-Inhibitors and Anti-CTLA-4 Antibodies

Rationale: Anti-CTLA-4 antibodies, e.g., ipilimumab, promote intratumoral T cell infiltration, but induce the upregulation of the inhibitory immune checkpoints VISTA and PD-L1 within the prostate TME at the same time [134]. Therefore, the simultaneous targeting of both immune checkpoints may help to overcome the adaptive mechanism of immune resistance.
A combination of the anti-CTLA-4 antibody tremelimumab plus PD-L1 inhibitor durvalumab was applied in men with chemotherapy-naïve mCRPC every 4 weeks (up to four doses), followed by durvalumab maintenance every 4 weeks (up to nine doses) in a single-arm pilot study (NCT03204812). Stable disease for >6 months was achieved in 24% of the patients. Median rPFS was 3.7 months (95% CI: 1.9 to 5.7), and median overall survival was 28.1 months (95% CI: 14.5 to 37.3). Of note, post-treatment evaluation of the bone microenvironment revealed transcriptional upregulation in myeloid and neutrophil immune subset signatures and increased the expression of inhibitory immune checkpoints indicating the development of immune resistance [135].
In the CheckMate650 trial, combinational therapy of ipilimumab and nivolumab was evaluated before (cohort A) and after docetaxel treatment (cohort B) in mCRPC (NCT02985957). ORR and disease control rates were 25% and 10%, and 46.9% and 13.3% in the pre- and post-chemotherapy setting, respectively. Two patients in each cohort had complete responses. An OS of 19.0 and 15.2 months was reported for cohort A and B. Initial biomarker analyses suggest improved activity in patients with high TMB (≥74.5 mutations/patient; ORR: 50% vs. 5.3%), evidence of a DDR (ORR: 36.4% vs. 23.1%), and increased PD-L1 expression (≥1%; ORR: 36.4 vs. 12.1%) [136].
A small phase 2 study (n = 15) also evaluated the efficacy of the combination of ipilimumab and nivolumab in patients with evidence of androgen receptor splice variant 7 (AR-V7) (NCT02601014). The rationale for this biomarker-based patient selection was the hypothesis that an increased rate of DNA repair mechanism defects may be present in AR-V7-positive patients. In fact, 40% of the patients carried DDR mutations and outcomes appeared generally better in DDR+ compared to DDR tumors with respect to PSA responses (33% vs. 0%; p = 0.14), with ORR (40% vs. 0%; p = 0.46) and PSA-PFS (HR 0.19; p = 0.11) [137].
In another nonrandomized phase 2 study, the combination of ipilimumab/nivolumab was examined in AR-V7-expressing mCRPC without (Cohort 1) or with (Cohort 2) the anti-androgen enzalutamide. For both groups only modest activity was reported without statistical differences between the cohorts. Lower alkaline phosphatase, lower circulating IL-7 and IL-6 levels, and higher circulating IL-17 levels were associated with improved OS [138].

3.3.2. PD-1/PD-L1 Antibodies and Androgen Receptor-Targeting Therapies

Rationale: Androgen receptor inhibitor enzalutamide is assumed to enhance IFNγ signaling and may sensitize tumor cells to immune-mediated cell killing, making it a candidate for combinations with PD-L1/PD-1 inhibitors. In addition, PD-L1 upregulation on dendritic cells in men with mCRPC either progressing on or refractory to enzalutamide has been reported.
In a phase 2 trial, enzalutamide was combined with PD-1 inhibitor pembrolizumab in mCRPC patients progressing on enzalutamide alone (NCT02312557). A PSA decline of ≥50% was reported in 18% of the patients, and 25% of the men with measurable disease at baseline achieved an objective response. Median OS for all patients was 21.9 months (95% CI: 14.7 to 28.4 months) and 41.7 months (95% CI: 22.16 to not reached (NR)) in men responding to the IC/NHA combinational therapy [139]. In addition, cohorts 4 and 5 of the Keynote199 trial (NCT02787005) evaluated pembrolizumab in combination with enzalutamide in chemotherapy-naive mCRPC patients after progression on enzalutamide therapy who had RECIST-measurable (cohort 4) or bone-predominant (cohort 5) disease. In cohort 4, ORR was reported in 12% of the patients. A DCR of 51% was observed for both cohorts. At the time of data cutoff, median OS was not reached in cohort 4 and was 19 months in cohort 5, respectively. Of note, the shorter median OS correlated with prior enzalutamide treatment <6 months [140]. Currently, combinational therapy of pembrolizumab and enzalutamide are compared to enzalutamide alone in phase 3 clinical trials in patients with hormone-sensitive (NCT04191096) or castration-resistant disease (NCT03834493) (Table 2).
In the IMbassador250 phase 3 clinical trial, a combination of PD-L1 inhibitor atezolizumab and enzalutamide was compared to enzalutamide alone in patients with mCRPC who had progressed on abiraterone and docetaxel or were not candidates for a taxane regimen (NCT03016312). The combination of atezolizumab and enzalutamid did not show an OS improvement (15.2 months (95% CI: 14.0, 17.0) vs. enzalutamide alone (16.6 months (95% CI: 14.7, 18.4), requiring early termination of the study [38,141]. In line with previous findings, low effector T cell and macrophage signatures, as well as reduced MHC class I and immune checkpoint signatures were observed in the majority of patients. In addition, markers of pre-existing immunity, such as PD-L1 IC expression ≥5% (IC2/3), CD8 T cell infiltration, and TMB ≥ 10 mutations per megabase, were rare. However, the presence of PD-L1 IC2/3 expression, high levels of CD8+ T cells, and established immune cell signatures were associated with longer PFS in the combination arm. Interestingly, an improved PFS was also found in patients with PTEN loss. This is of such interest because loss of PTEN activity is known to be associated with an immunosuppressive milieu (see above). The addition of atezolizumab could potentially reverse this [38].

3.3.3. PD-1/PD-L1 and Chemotherapy

Rationale: Cytotoxic cell death and subsequent antigen release provides immune stimulation, common to conventional and targeted anticancer agents. However, in the past years, there is emerging evidence that the efficacy of chemotherapy does not only involve direct cytostatic/cytotoxic effects, but also relies on the (re)activation of tumor-targeting immune responses [142]. Indeed, a reduction of circulating Treg cells and MDSCs has been described associated with a more immunopermissive TME.
In cohort B of CheckMate9KD, combinational therapy of nivolumab and docetaxel was evaluated in mCRPC patients, who had previously received up to two NHA. An ORR and PSA reduction >50% from baseline was reported for 40% (95% CI, 25.7–55.7) and 46.9% (95% CI, 35.7–58.3) of the patients, respectively. rPFS and OS were 9.0 months (95% CI, 8.0–11.6) and 18.2 months (95% CI, 14.6–20.7), respectively [143,144]. In subpopulations with versus without prior NHA, the ORR was 38.7% versus 42.9% and the PSA reduction >50% was 39.6% vs. 60.7%. In addition, median rPFS was improved from 8.5 to 12.0 months and median OS from 16.2 months versus not reached. Of note, preliminary biomarker analyses revealed no clear association between HRD or TMB with tumor reduction or with decrease in PSA from baseline. In addition, CDK12 mutations did not correlate with response or PSA reduction.
The additional benefit of adding nivolumab or pembrolizumab to chemotherapy with docetaxel are currently being determined in two randomized phase 3 trials (NCT04100018; NCT03834506; Table 2).

3.3.4. PD-1/PD-L1 and PARP Inhibitors

Rationale: HRR defects have been associated with improved response to immunotherapy in PCa (see above). PARP inhibition may potentiate DNA damage and inefficient repair in tumors, and thus may cause immunologically relevant mutations [145].
In a phase 1/2 clinical trial, durvalumab was evaluated in combination with PARP inhibitor (PARPi) olaparib in patients with mCRPC with and without somatic or germline DDR mutations. The median rPFS of patients with alterations in DDR genes was 16.1 months (95% CI: 7.8–18.1 months) with a 12-month PFS probability of 83.3% (95% CI: 27.3– 94.5%) compared with a 12-month probability of 36.4% (95% CI: 11.2–62.7%) for those without mutations; p = 0.031). Remarkably, patients’ baseline fraction of MDSCs correlated with response to therapy revealed by a prolonged PFS of those whose percentage of MDSCs among total viable cells was below the median baseline (p = 0.041) [146].
Cohort A1 of CheckMate9KD received nivolumab in combination with PARPi rucaparib after 1–2 prior taxane-based chemotherapy regimens and up to 2 NHA for mCRPC (NCT03338790). In HRD+ patients, nivolumab and rucaparib achieved a confirmed ORR and a PSA reduction of ≥50% from baseline in 17.2% (95% CI, 5.8–35.8) and 18.2% (95% CI, 8.2–32.7), respectively. In contrast, patients with HRD tumors did not appear to benefit from either drug [147].
In Cohort A2 of CheckMate9KD, the combination of nivolumab and rucaparib was applied to mCRPC patients without previous taxane treatment and 1–2 prior NHA (NCT03338790). In the pre-chemotherapy setting, confirmed ORR was reported for 25% (95% CI, 8.7–49.1) and a confirmed PSAresponse for 41.9 (95% CI, 24.5–60.9) of HRD + patients, respectively. Remarkably, a PSA reduction of >50% from baseline was found in 84.6% of the patients. Again, clinical activity in patients with HRD tumors was limited [148].
An ongoing randomized phase 3 clinical trial is currently evaluating combinational therapy of pembrolizumab and olaparib compared to the second NHA in HRD-unselected mCRPC patients (NCT03834519; Table 2).

3.3.5. PD-1/PD-L1-Inhibitors and Tyrosinkinase Inhibitors (TKI)

Rationale: Cabozantinib inhibits several tyrosine kinases including MET, VEGF receptors, and TAM family of kinases (TYRO3, MER, and AXL) [149]. Interestingly, it promotes an immune-permissive environment that may enhance response to immune checkpoint inhibitors [150,151].
On the annual meeting of the European Society of Medical Oncology (ESMO) 2021, data of the COSMIC-021 trial cohort 6 were presented, examining the role of cabozantinib and atezolizumab in patients with mCRPC [152]. Pretreatment with at least one NHA was a prerequisite for study inclusion, while chemotherapy was only allowed in the hormone-sensitive setting. At baseline, 77% of the patients had measurable visceral metastases (32%) or extrapelvic lymphadenopathy (EPLN) (60%). ORR per investigator was 23% for all patients and 27% for patients with visceral or EPLN with 2% CR. Remarkably, a disease control rate of 84% and 88% was reported for both groups, respectively. The ongoing phase 3 trial CONTACT-02 evaluates the TKI/CI combination in mCRPC patients with measurable disease who have been pre-treated with one NHA (NCT04446117; Table 2).
In addition, multikinase inhibitor lenvatinib is investigated in combination with pembrolizumab in neuroendocrine PCa (NCT04848337; Table 2).

3.3.6. PD-1/PD-L1 Inhibitors and Radiotherapeutic Approaches

Rationale: Abscopal effects with partial or complete eradication of tumors distant from the local radiation fields have been observed in melanoma and lung cancer patients receiving immunotherapy and local radiotherapy [153,154].
In a phase 1b study, PD-L1 inhibitor atezolizumab was combined with radium-223 applying three different treatment schedules (NCT02814669). However, no clear evidence of additional clinical benefit was observed in mCRPC patients with bone and lymph node and/or visceral metastases independent of the regiment used. Of note, the combination was accompanied by increased toxicity compared to either drug alone [155]. Currently, radium-223 is evaluated in combination with nivolumab (NCT04109729), pembrolizumab (NCT03093428), or avelumab and radiation-enhancing medication M3814 (NCT04071236) (Table 2).
Current clinical trials are evaluating the combination of CI with PSMA ligand therapy or metastases-directed therapy (NCT05150236; Table 2). Recently, interim results of the PRINCE trial on the combination of pembrolizumab and radioligand therapy with 177-Lu-PSMA were presented. Here, 37 patients received an average of four doses of 177-Lu-PSMA-617 and eight doses of pembrolizumab. A PSA decline of at least 50% was observed in 73% of patients, and 78% of patients achieved partial remission according to RECIST 1.1. The toxicity profile was similar to that of the single agents. Further results are eagerly awaited.

3.4. Bispecific T Cell Engagers

Bispecific T cell engagers (BiTEs) are synthetic proteins designed to activate and target T cells to tumor cells (Figure 2). The structure of BiTEs is based on variable antibody fragments. A BiTE is made up of two different specific binding domains, and each binding domain is formed by two single-chain variable fragments connected by a linker. In general, one of the two domains is specific for CD3, a cell surface marker of T cells that is required for co-stimulation in the T cell receptor complex. The second domain can be specifically adapted to the tumor antigen of interest [156]. Upon binding of the BiTE to the T cell and tumor cell, an immunologic synapse is formed and cytotoxic T cells initiate tumor cell lysis, without the need for further co-stimulation. This is of high value in tumors where MHC class I expression is downregulated, and thus is a promising strategy for PCa [157]. At the immunological synapse perforin and granzymes are released by cytotoxic T cells and eventually cause tumor cell death. As a result of activation, the T cells proliferate, thereby potentiating the antitumor effects [156]. BiTEs are designed to have a higher affinity to tumor-specific targets than to CD3 to reduce binding of T cells in the absence of tumor cells. A reduction of CD3 affinity also decreases cytokine release and, consequently also side effects [158].
Currently, conventional BiTEs must be administered as continuous infusions due to their short half-life. For this reason, extending the half-life is an important scientific challenge to improve clinical applicability. One method to increase half-life time in circulation is the fusion of BiTEs with Fc fragments. For instance, AMG160 is an anti-PSMA half-life extended BiTE that has successfully been tested in animal models and provided first promising clinical results in PCa [159,160]. Thus, at the ESMO annual meeting 2020, preliminary results of an ongoing phase 1 trial were presented (NCT03792841) [160]. At the time of data cutoff, 43 patients had received at least one dose of AMG160 and 44.2% of the patients remained on treatment. However, 95.3% of the patients experienced treatment-related adverse events (TRAE) with three reversible dose-limiting toxicities. Of note, none resulted in treatment discontinuation and no grade 5 events were reported. Confirmed and additional unconfirmed PSA responses (≥30% decrease) were achieved by 27.6% and 11.4% of the patients, respectively. In 23.1% of the men, previously detectable circulating tumor cells disappeared during the course of therapy. Confirmed and unconfirmed responses as well as stable disease according to RECIST1.1 were observed in 13.3%, 6.7%, and 53.3% of the study participants, respectively. A phase 3 trial is currently in preparation. Another approach to overcome the limited half-life time is the use of an injectable polymer depot. Anti-PSMA-BiTEs enclosed in a biopolymer are released as the biopolymer gets slowly degraded. In mouse xenograft models of PCa, the biopolymer showed a low inflammatory potential and the BiTE depots effectively maintained BiTE plasma concentration. Especially in tumors with low PSMA expression, the inhibition of tumor growth was improved with the use of a BiTE depot compared to daily injection of the BiTE alone [159].
An intensive search for alternative target antigens is currently underway, the first of which are now being evaluated in the preclinical setting and early clinical trials. Thus, a BiTE targeting Glypican-1, a heparan sulfate proteoglycan that is overexpressed in PCa with a correlation to the Gleason score, has been designed on the basis of the CD3 binding sequence of blinatumomab in a standard BiTE format. Promising preclinical results have been reported including T cell activation and cytokine release [161]. In addition, a BiTE targeting disintegrin and metalloproteinase 17 (ADAM17), a transmembrane protease, and an anti-ADAM17 BiTEs-mediated specific lysis of ADAM17-expressing cells including PCa cell lines have been analyzed [162]. Prostate stem cell antigen (PSCA), a glycosylphosphatidylinositol (GPI)-anchored cell surface protein, upregulated in different malignancies including mCRPC, is serving as a target for GEM3PSCA, an affinity-tailored T cell adaptor, currently evaluated in PSCA-positive PCa in a phase 1 clinical trial (NCT03927573; Table 3).
AMG-757 is another half-life extended BiTE targeted against delta-like ligand 3 (DLL3), a notch ligand involved in neuroendocrine differentiation. While AMG-757 has been successfully tested in preclinical models of small cell lung cancer, it might as well be effective against neuroendocrine PCa as DLL3 is also upregulated in these tumors [163,164]. A phase 1 clinical trial is carried out in patients with de novo or treatment emergent neuroendocrine prostate cancer (NCT04702737; Table 3).
A major downside of BiTEs, however, is the activation of immune checkpoint molecules, such as PD-1 or LAG-3, as a consequence of T cell activation. Therefore, combination of BiTEs with immune checkpoint inhibitors might be able to overcome treatment resistance [165]. A recent study in an animal model has revealed that especially immunologically cold tumor with low T cell infiltration may benefit from combination of BiTE therapy and concurrent immune checkpoint inhibition [166].

3.5. Chimeric Antigen Receptor T Cells (CAR-T Cells)

Chimeric antigen receptor T cells (CAR-T Cells) are genetically modified T cells that are transfected with a chimeric antigen receptor directed against a tumor antigen (Figure 2). Following in vitro expansion, the CAR-T cells are transfused back into the patient. Optimization of CAR-T constructs mainly focuses on the intracellular signal transduction domains. Common to all is the CD3-zeta signaling domain, while further costimulatory domains have been added in the newer generations to enhance survival and proliferation of CAR-T cells [167,168]. CAR-T cell therapy has already profoundly improved treatment options in adult lymphoblastic leukemia of B cell lineage and of Non-Hodgkin B cell lymphoma as well as multiple myeloma [169,170,171]. In comparison to conventional cytotoxic chemotherapy or immunotherapy with monoclonal antibodies, CAR-T cell therapy has been found capable to induce durable complete responses after a single treatment course. This is based upon the ability of CAR-T cells to expand in vivo and to persist for several years, which leads to continuous therapeutic efficacy and tumor control [168,172].
For CAR-T-based therapy of metastatic PCa, several antigens and different CAR-T constructs are currently under clinical investigation (Table 3). PSMA has been identified as an attractive target for CAR-T cell therapy due to its consistent membranous expression and because the majority of mCRPC are positive for PSMA [173]. Of note, PSMA expression has been also described in small intestine, kidney, central nervous system, and salivary glands [174,175]. Therefore, the possible on-target off-tumor toxicity of PSMA-directed CAR-T therapy with regard to these tissues has to be accounted for. Other targets currently under investigation are PSCA and kallikrein 2 (KLK2), both reported with high expression in prostate cancer [176,177,178] (Table 3). Challenges in CAR-T cell therapy of solid tumors and prostate cancer in particular include the previously described immunosuppressive tumor environment as well as reduced homing and decreased persistence of CAR-T cells [179,180]. To counter these barriers, several approaches have been proposed. However, to date only few clinical results are available, mainly reported in press releases or congress abstracts.
In a phase 1 trial, CAR-T cells with specificity to PSMA were co-administered with IL-2 after a non-myeloablative chemotherapy with cyclophosphamide and fludarabine. All toxicities observed were attributed to chemotherapy or IL-2 treatment. Remarkably, despite the early phase 1 study design, two of five patients displayed a PSA response [181]. Another attempt to increase the effectiveness of CAR-T cells is the co-expression of a dominant negative TGF-β receptor in CAR-T cells directed to PSMA. The resulting decrease of the immunosuppressive signaling of TGF beta led to the enhancement of CAR-T proliferation, antitumor activity, and persistence in a pre-clinical model of aggressive PCa [64,182]. Consequently, “augmented” CAR-T cells with a dominant negative TGF-β receptor were evaluated in a phase 1 clinical trial (NCT 03089203). Remarkably, cytokin-release syndrom has been observed as a common toxicity in this trial. This suggests that CAR-T cells were able to withstand the immunosuppressive tumor microenvironment and were stimulated to proliferate by successful binding to the PSMA antigen. Unfortunately, on higher CAR-T doses in this study, excessive toxicity was observed, which led to fatal outcomes in one patient due to neurotoxicity and macrophage activation syndrome. Of note, at this very early stage PSA response was observed in three of six patients [183].
Several CAR-T constructs have been developed that harbor safety switches to improve mitigation of toxicities. This includes “on-switches”, which require the presence of a small molecule to enable CAR-T cell activation. For instance, the complete CAR can be formed by two different subunits, one containing only the antigen-binding domain and the other harboring the signal transduction domain within the T cell. Only after infusion of the antigen-binding domain do both subunits dimerize and mediate T cell activation and tumor cell lysis. Withdrawal of the antigen-binding site infusion will disrupt CAR-T cell activation. This concept is currently investigated by the UC02-PSMA trial, which is using a “treatment module” adapter molecule that binds to the target antigen PSMA and the CAR-T cell, which itself is not able to bind to PSMA. The treatment module is given as continuous infusion due to its short half-life. Stopping of the treatment module infusion is expected to rapidly counter CAR-T-mediated toxicities. By using different adaptor molecules with the same CAR epitope, one CAR-T cell population can target multiple tumor-associated antigens [184].
As alternative measure for safety, “off-switches” are used to stop CAR signaling and CAR-T cell proliferation. This includes small molecule inhibitors of T-cell receptor signaling as well as the induction of CAR degradation [184]. Kill switches are permanent off-switches that cause apoptosis of the CAR-T cells upon administration of a small molecule. One example is the transduction of CAR-T cells with a suicide gene, such as the Herpes simplex virus thymidine kinase, a mechanism that has so far been used in hematopoietic stem cell transplantations [185]. Additionally, small molecule-mediated dimerization of a transgenic caspase like rimiducid can be applied for CAR-T cell killing [168,186].
Another promising and noteworthy approach of CAR-T therapy is the use of allogeneic rather than the standard autologous T cells for CAR-T manufacturing [187,188]. Besides the logistical advantages of allogeneic CAR-T production, CAR-T cells manufactured from the T cells of healthy donors may provide superior immunological properties and eventually improved efficacy [189]. Although several clinical trials are currently investigating allogeneic CAR-T therapy in hematological und solid malignancies, to the best of our knowledge this is not yet the case for PCa.

4. Future Directions

The start of immunotherapy in the treatment of advanced PCa was bumpy. Many of the high expectations could not be met at first. The growing knowledge about the specific immunosuppressive milieu of PCa and possible counter-regulatory interventions give hope that PCa patients will also be able to benefit from immunotherapy in the future. Various combination therapies to improve the effectiveness of checkpoint inhibitors are currently underway and their results are eagerly awaited.
A perceived setback was the phase 3 results on atezolizumab and enzalutamide showing no OS benefit for the “‘intention to treat” population, and thus requiring early termination of the trial. Nonetheless, important findings on potential biomarkers show that there are PCa patients who will benefit from immunotherapy [38]. In addition, long-term analyses of ipilimumab have recently demonstrated two to three times higher survival rates at 3 years, despite an initially negative result of the trial. Great hope rests on new treatment strategies such as BiTEs or CAR-T cells. However, the recent attempts to counteract immunosuppressive factors by additional genetic modification of CAR-T cells have led to unexpectedly severe toxicities in addition to improved efficacy, which is to be resolved in the future. In conclusion, despite many relevant questions that remain to be addressed, the intensive scientific efforts on different levels give hope that there is a light at the end of the tunnel.

Author Contributions

Conceptualization, G.v.A., L.M. and S.A.D.; resources, all authors; writing—original draft preparation, all authors; writing—review and editing, G.v.A. and P.K., visualization, G.v.A. and L.M.; supervision, G.v.A. All authors have read and agreed to the published version of the manuscript.

Funding

The research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

G.v.A.: Consulting or expert activities Advisory Boards: Roche, BMS, Astellas, Sanofi, Janssen, MSD, Merck Serono, Pfizer, Bayer, Astra Zeneca, EISAI; Honoraria/lectures/travel expenses/congress support: Roche, BMS, Astellas, Sanofi, Ipsen, EISAI, Pierre Fabre, MSD, Astra Zeneca, Janssen, Bayer, Merck Serono, Pfizer; Funding of scientific studies in the context of industry-sponsored studies: Roche, BMS, MSD, Astra Zeneca, Sanofi, AvenCell, Pfizer, Exelixis, Amgen. W.A.: Funding of scientific studies: Biontech; Advisory board fees: Janssen; congress support: Biontech, Janssen. P.K.: Advisory board fees: Janssen-Cliag; Stockholder: Abbvie. C.B.: Advisory board fees: Sanofi Aventis, Merck KgA, Bristol-Myers Squibb, Merck Sharp and Dohme, Lilly Imclone, Bayer Healthcare, GSO Contract Research, AOK Rheinland-Hamburg, Novartis. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Abbreviations

CTLA-4cytotoxic T lymphocyte antigen 4
PD-1programmed death-1
PD-L1PD-1 ligand
PCaprostate cancer
TMEtumor microenvironment
BiTEbispecific T cell engager
CAR-T cellschimeric antigen receptor T cells
TMИtumor mutational burden
MHCmajor histocompatibility complex
CTLcytotoxic T lymphocytes
TILstumor-infiltrating lymphocytes
DDRDNA damage repair
MMRmismatch repair
MSImicrosatellite instability
HRRhomologous recombination repair
Tregregulatory T cells
AVPCaggressive variant of prostate cancer
MDSCsmyeloid-derived suppressor cells
AKandrogen receptor
ADTandrogen-deprivation therapy
TAMtumor-associated macrophages
TNF-α tumor necrosis factor-α
MSCs mesenchymal stromal cells
CAFscancer-associated fibroblasts
EMTepithelial–mesenchymal transition
PAPprostatic acid phosphatase
ATPadenosine triphosphate
DAMPdanger-associated molecular pattern
TAAtumor-associated antigens
PSAprostate-specific antigen
PSMAprostate-specific membrane antigen
PSCAprostate stem cell antigen
APCantigen-presenting cell
DCdendritic cell
OSoverall survival
mCRPCmetastatic castration-resistant prostate cancer
PBMCsperipheral blood mononuclear cells
mHNPCmetastatic hormone-naïve prostate cancer
PPVpersonalized peptide vaccination
PFSprogression-free survival
CIimmune checkpoint inhibitor
ORR overall response rate
SDstable disease
DCRdisease control rate
PR partial response
HRDhomologous recombination deficiency
TRAETreatment-related adverse events

References

  1. Siegel, R.L.; Miller, K.D.; Jemal, A. Cancer statistics, 2020. CA Cancer J. Clin. 2020, 70, 7–30. [Google Scholar] [CrossRef] [PubMed]
  2. Ferro, M.; Lucarelli, G.; Crocetto, F.; Dolce, P.; Verde, A.; La Civita, E.; Zappavigna, S.; de Cobelli, O.; Di Lorenzo, G.; Facchini, B.A.; et al. First-line systemic therapy for metastatic castration-sensitive prostate cancer: An updated systematic review with novel findings. Crit. Rev. Oncol. Hematol. 2021, 157, 103198. [Google Scholar] [CrossRef] [PubMed]
  3. Smith, M.R.; Hussain, M.; Saad, F.; Fizazi, K.; Sternberg, C.N.; Crawford, E.D.; Kopyltsov, E.; Park, C.H.; Alekseev, B.; Montesa-Pino, Á.; et al. Darolutamide and Survival in Metastatic, Hormone-Sensitive Prostate Cancer. N. Engl. J. Med. 2022, in press. [Google Scholar] [CrossRef] [PubMed]
  4. Fizazi, K.; Maldonado, X.; Foulon, S.; Roubaud, G.; McDermott, R.S.; Flechon, A.; Tombal, B.F.; Supiot, S.; Berthold, D.R.; Ronchin, P.; et al. A phase 3 trial with a 2x2 factorial design of abiraterone acetate plus prednisone and/or local radiotherapy in men with de novo metastatic castration-sensitive prostate cancer (mCSPC): First results of PEACE-1. J. Clin. Oncol. 2021, 39, 5000. [Google Scholar] [CrossRef]
  5. Hofman, M.S.; Emmett, L.; Sandhu, S.; Iravani, A.; Joshua, A.M.; Goh, J.C.; Pattison, D.A.; Tan, T.H.; Kirkwood, I.D.; Ng, S.; et al. [177Lu]Lu-PSMA-617 versus cabazitaxel in patients with metastatic castration-resistant prostate cancer (TheraP): A randomised, open-label, phase 2 trial. Lancet 2021, 397, 797–804. [Google Scholar] [CrossRef]
  6. De Wit, R.; de Bono, J.; Sternberg, C.N.; Fizazi, K.; Tombal, B.; Wulfing, C.; Kramer, G.; Eymard, J.C.; Bamias, A.; Carles, J.; et al. Cabazitaxel versus Abiraterone or Enzalutamide in Metastatic Prostate Cancer. N. Engl. J. Med. 2019, 381, 2506–2518. [Google Scholar] [CrossRef]
  7. Sartor, O.; de Bono, J.; Chi, K.N.; Fizazi, K.; Herrmann, K.; Rahbar, K.; Tagawa, S.T.; Nordquist, L.T.; Vaishampayan, N.; El-Haddad, G.; et al. Lutetium-177–PSMA-617 for Metastatic Castration-Resistant Prostate Cancer. N. Engl. J. Med. 2021, 385, 1091–1103. [Google Scholar] [CrossRef]
  8. De Bono, J.; Mateo, J.; Fizazi, K.; Saad, F.; Shore, N.; Sandhu, S.; Chi, K.N.; Sartor, O.; Agarwal, N.; Olmos, D.; et al. Olaparib for Metastatic Castration-Resistant Prostate Cancer. N. Engl. J. Med. 2020, 382, 2091–2102. [Google Scholar] [CrossRef]
  9. Alatrash, G.; Jakher, H.; Stafford, P.D.; Mittendorf, E.A. Cancer immunotherapies, their safety and toxicity. Expert Opin. Drug Saf. 2013, 12, 631–645. [Google Scholar] [CrossRef]
  10. Wolchok, J.D.; Chiarion-Sileni, V.; Gonzalez, R.; Rutkowski, P.; Grob, J.J.; Cowey, C.L.; Lao, C.D.; Wagstaff, J.; Schadendorf, D.; Ferrucci, P.F.; et al. Overall Survival with Combined Nivolumab and Ipilimumab in Advanced Melanoma. N. Engl. J. Med. 2017, 377, 1345–1356. [Google Scholar] [CrossRef]
  11. Motzer, R.J.; Tannir, N.M.; McDermott, D.F.; Arén Frontera, O.; Melichar, B.; Choueiri, T.K.; Plimack, E.R.; Barthélémy, P.; Porta, C.; George, S.; et al. Nivolumab plus Ipilimumab versus Sunitinib in Advanced Renal-Cell Carcinoma. N. Engl. J. Med. 2018, 378, 1277–1290. [Google Scholar] [CrossRef]
  12. Shields, M.D.; Marin-Acevedo, J.A.; Pellini, B. Immunotherapy for Advanced Non–Small Cell Lung Cancer: A Decade of Progress. Am. Soc. Clin. Oncol. Educ. Book 2021, 41, e105–e127. [Google Scholar] [CrossRef]
  13. Kantoff, P.W.; Higano, C.S.; Shore, N.D.; Berger, E.R.; Small, E.J.; Penson, D.F.; Redfern, C.H.; Ferrari, A.C.; Dreicer, R.; Sims, R.B.; et al. Sipuleucel-T immunotherapy for castration-resistant prostate cancer. N. Engl. J. Med. 2010, 363, 411–422. [Google Scholar] [CrossRef] [Green Version]
  14. Klempner, S.J.; Fabrizio, D.; Bane, S.; Reinhart, M.; Peoples, T.; Ali, S.M.; Sokol, E.S.; Frampton, G.; Schrock, A.B.; Anhorn, R.; et al. Tumor Mutational Burden as a Predictive Biomarker for Response to Immune Checkpoint Inhibitors: A Review of Current Evidence. Oncologist 2020, 25, e147–e159. [Google Scholar] [CrossRef] [Green Version]
  15. Lawrence, M.S.; Stojanov, P.; Polak, P.; Kryukov, G.V.; Cibulskis, K.; Sivachenko, A.; Carter, S.L.; Stewart, C.; Mermel, C.H.; Roberts, S.A.; et al. Mutational heterogeneity in cancer and the search for new cancer-associated genes. Nature 2013, 499, 214–218. [Google Scholar] [CrossRef]
  16. Ryan, M.J.; Bose, R. Genomic Alteration Burden in Advanced Prostate Cancer and Therapeutic Implications. Front. Oncol. 2019, 9, 1287. [Google Scholar] [CrossRef] [Green Version]
  17. Wagle, M.C.; Castillo, J.; Srinivasan, S.; Holcomb, T.; Yuen, K.C.; Kadel, E.E.; Mariathasan, S.; Halligan, D.L.; Carr, A.R.; Bylesjo, M.; et al. Tumor Fusion Burden as a Hallmark of Immune Infiltration in Prostate Cancer. Cancer Immunol. Res. 2020, 8, 844–850. [Google Scholar] [CrossRef] [Green Version]
  18. Haffner, M.C.; Guner, G.; Taheri, D.; Netto, G.J.; Palsgrove, D.N.; Zheng, Q.; Guedes, L.B.; Kim, K.; Tsai, H.; Esopi, D.M.; et al. Comprehensive Evaluation of Programmed Death-Ligand 1 Expression in Primary and Metastatic Prostate Cancer. Am. J. Pathol. 2018, 188, 1478–1485. [Google Scholar] [CrossRef] [Green Version]
  19. Palicelli, A.; Bonacini, M.; Croci, S.; Magi-Galluzzi, C.; Cañete-Portillo, S.; Chaux, A.; Bisagni, A.; Zanetti, E.; De Biase, D.; Melli, B.; et al. What Do We Have to Know about PD-L1 Expression in Prostate Cancer? A Systematic Literature Review. Part 1: Focus on Immunohistochemical Results with Discussion of Pre-Analytical and Interpretation Variables. Cells 2021, 10, 3166. [Google Scholar] [CrossRef]
  20. Ebelt, K.; Babaryka, G.; Frankenberger, B.; Stief, C.G.; Eisenmenger, W.; Kirchner, T.; Schendel, D.J.; Noessner, E. Prostate cancer lesions are surrounded by FOXP3+, PD-1+ and B7-H1+ lymphocyte clusters. Eur. J. Cancer 2009, 45, 1664–1672. [Google Scholar] [CrossRef]
  21. Gevensleben, H.; Dietrich, D.; Golletz, C.; Steiner, S.; Jung, M.; Thiesler, T.; Majores, M.; Stein, J.; Uhl, B.; Müller, S.; et al. The Immune Checkpoint Regulator PD-L1 Is Highly Expressed in Aggressive Primary Prostate Cancer. Clin. Cancer Res. 2016, 22, 1969–1977. [Google Scholar] [CrossRef] [Green Version]
  22. Baas, W.; Gershburg, S.; Dynda, D.; Delfino, K.; Robinson, K.; Nie, D.; Yearley, J.H.; Alanee, S. Immune Characterization of the Programmed Death Receptor Pathway in High Risk Prostate Cancer. Clin. Genitourin Cancer 2017, 15, 577–581. [Google Scholar] [CrossRef] [PubMed]
  23. Calagua, C.; Russo, J.; Sun, Y.; Schaefer, R.; Lis, R.; Zhang, Z.; Mahoney, K.; Bubley, G.J.; Loda, M.; Taplin, M.E.; et al. Expression of PD-L1 in Hormone-naïve and Treated Prostate Cancer Patients Receiving Neoadjuvant Abiraterone Acetate plus Prednisone and Leuprolide. Clin. Cancer Res. 2017, 23, 6812–6822. [Google Scholar] [CrossRef] [Green Version]
  24. Ness, N.; Andersen, S.; Khanehkenari, M.R.; Nordbakken, C.V.; Valkov, A.; Paulsen, E.E.; Nordby, Y.; Bremnes, R.M.; Donnem, T.; Busund, L.T.; et al. The prognostic role of immune checkpoint markers programmed cell death protein 1 (PD-1) and programmed death ligand 1 (PD-L1) in a large, multicenter prostate cancer cohort. Oncotarget 2017, 8, 26789–26801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Fankhauser, C.D.; Schüffler, P.J.; Gillessen, S.; Omlin, A.; Rupp, N.J.; Rueschoff, J.H.; Hermanns, T.; Poyet, C.; Sulser, T.; Moch, H.; et al. Comprehensive immunohistochemical analysis of PD-L1 shows scarce expression in castration-resistant prostate cancer. Oncotarget 2018, 9, 10284–10293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Palicelli, A.; Croci, S.; Bisagni, A.; Zanetti, E.; De Biase, D.; Melli, B.; Sanguedolce, F.; Ragazzi, M.; Zanelli, M.; Chaux, A.; et al. What Do We Have to Know about PD-L1 Expression in Prostate Cancer? A Systematic Literature Review (Part 6): Correlation of PD-L1 Expression with the Status of Mismatch Repair System, BRCA, PTEN, and Other Genes. Biomedicines 2022, 10, 236. [Google Scholar] [CrossRef]
  27. Petitprez, F.; Fossati, N.; Vano, Y.; Freschi, M.; Becht, E.; Lucianò, R.; Calderaro, J.; Guédet, T.; Lacroix, L.; Rancoita, P.M.V.; et al. PD-L1 Expression and CD8(+) T-cell Infiltrate are Associated with Clinical Progression in Patients with Node-positive Prostate Cancer. Eur. Urol. Focus 2019, 5, 192–196. [Google Scholar] [CrossRef] [PubMed]
  28. Sharma, M.; Yang, Z.; Miyamoto, H. Immunohistochemistry of immune checkpoint markers PD-1 and PD-L1 in prostate cancer. Medicine 2019, 98, e17257. [Google Scholar] [CrossRef] [PubMed]
  29. Abida, W.; Cheng, M.L.; Armenia, J.; Middha, S.; Autio, K.A.; Vargas, H.A.; Rathkopf, D.; Morris, M.J.; Danila, D.C.; Slovin, S.F.; et al. Analysis of the Prevalence of Microsatellite Instability in Prostate Cancer and Response to Immune Checkpoint Blockade. JAMA Oncol. 2019, 5, 471–478. [Google Scholar] [CrossRef]
  30. Pritchard, C.C.; Morrissey, C.; Kumar, A.; Zhang, X.; Smith, C.; Coleman, I.; Salipante, S.J.; Milbank, J.; Yu, M.; Grady, W.M.; et al. Complex MSH2 and MSH6 mutations in hypermutated microsatellite unstable advanced prostate cancer. Nat. Commun. 2014, 5, 4988. [Google Scholar] [CrossRef] [Green Version]
  31. Linch, M.; Goh, G.; Hiley, C.; Shanmugabavan, Y.; McGranahan, N.; Rowan, A.; Wong, Y.N.S.; King, H.; Furness, A.; Freeman, A.; et al. Intratumoural evolutionary landscape of high-risk prostate cancer: The PROGENY study of genomic and immune parameters. Ann. Oncol. 2017, 28, 2472–2480. [Google Scholar] [CrossRef]
  32. Le, D.T.; Durham, J.N.; Smith, K.N.; Wang, H.; Bartlett, B.R.; Aulakh, L.K.; Lu, S.; Kemberling, H.; Wilt, C.; Luber, B.S.; et al. Mismatch repair deficiency predicts response of solid tumors to PD-1 blockade. Science 2017, 357, 409–413. [Google Scholar] [CrossRef] [Green Version]
  33. Nava Rodrigues, D.; Rescigno, P.; Liu, D.; Yuan, W.; Carreira, S.; Lambros, M.B.; Seed, G.; Mateo, J.; Riisnaes, R.; Mullane, S.; et al. Immunogenomic analyses associate immunological alterations with mismatch repair defects in prostate cancer. J. Clin. Investig. 2018, 128, 4441–4453. [Google Scholar] [CrossRef] [Green Version]
  34. Abida, W.; Armenia, J.; Gopalan, A.; Brennan, R.; Walsh, M.; Barron, D.; Danila, D.; Rathkopf, D.; Morris, M.; Slovin, S.; et al. Prospective Genomic Profiling of Prostate Cancer Across Disease States Reveals Germline and Somatic Alterations That May Affect Clinical Decision Making. JCO Precis. Oncol. 2017, 2017, 1–16. [Google Scholar] [CrossRef]
  35. Jenzer, M.; Kess, P.; Nientiedt, C.; Endris, V.; Kippenberger, M.; Leichsenring, J.; Stogbauer, F.; Haimes, J.; Mishkin, S.; Kudlow, B.; et al. The BRCA2 mutation status shapes the immune phenotype of prostate cancer. Cancer Immunol. Immunother. 2019, 68, 1621–1633. [Google Scholar] [CrossRef] [Green Version]
  36. Heijink, A.M.; Talens, F.; Jae, L.T.; van Gijn, S.E.; Fehrmann, R.S.N.; Brummelkamp, T.R.; van Vugt, M.A.T.M. BRCA2 deficiency instigates cGAS-mediated inflammatory signaling and confers sensitivity to tumor necrosis factor-alpha-mediated cytotoxicity. Nat. Commun. 2019, 10, 100. [Google Scholar] [CrossRef] [Green Version]
  37. Powles, T.; Loriot, Y.; Ravaud, A.; Vogelzang, N.J.; Duran, I.; Retz, M.; Giorgi, U.D.; Oudard, S.; Bamias, A.; Koeppen, H.; et al. Atezolizumab (atezo) vs. chemotherapy (chemo) in platinum-treated locally advanced or metastatic urothelial carcinoma (mUC): Immune biomarkers, tumor mutational burden (TMB), and clinical outcomes from the phase III IMvigor211 study. J. Clin. Oncol. 2018, 36, 409. [Google Scholar] [CrossRef]
  38. Powles, T.; Yuen, K.C.; Gillessen, S.; Kadel, E.E.; Rathkopf, D.; Matsubara, N.; Drake, C.G.; Fizazi, K.; Piulats, J.M.; Wysocki, P.J.; et al. Atezolizumab with enzalutamide versus enzalutamide alone in metastatic castration-resistant prostate cancer: A randomized phase 3 trial. Nat. Med. 2022, 28, 144–153. [Google Scholar] [CrossRef]
  39. Aparicio, A.M.; Shen, L.; Tapia, E.L.; Lu, J.F.; Chen, H.C.; Zhang, J.; Wu, G.; Wang, X.; Troncoso, P.; Corn, P.; et al. Combined Tumor Suppressor Defects Characterize Clinically Defined Aggressive Variant Prostate Cancers. Clin. Cancer Res. 2016, 22, 1520–1530. [Google Scholar] [CrossRef] [Green Version]
  40. Hamid, A.A.; Gray, K.P.; Shaw, G.; MacConaill, L.E.; Evan, C.; Bernard, B.; Loda, M.; Corcoran, N.M.; Van Allen, E.M.; Choudhury, A.D.; et al. Compound Genomic Alterations of TP53, PTEN, and RB1 Tumor Suppressors in Localized and Metastatic Prostate Cancer. Eur. Urol. 2019, 76, 89–97. [Google Scholar] [CrossRef]
  41. Robinson, D.; Van Allen, E.M.; Wu, Y.M.; Schultz, N.; Lonigro, R.J.; Mosquera, J.M.; Montgomery, B.; Taplin, M.E.; Pritchard, C.C.; Attard, G.; et al. Integrative clinical genomics of advanced prostate cancer. Cell 2015, 161, 1215–1228. [Google Scholar] [CrossRef] [Green Version]
  42. Peng, W.; Chen, J.Q.; Liu, C.; Malu, S.; Creasy, C.; Tetzlaff, M.T.; Xu, C.; McKenzie, J.A.; Zhang, C.; Liang, X.; et al. Loss of PTEN Promotes Resistance to T Cell-Mediated Immunotherapy. Cancer Discov. 2016, 6, 202–216. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Cao, Y.; Wang, H.; Yang, L.; Zhang, Z.; Li, C.; Yuan, X.; Bu, L.; Chen, L.; Chen, Y.; Li, C.M.; et al. PTEN-L promotes type I interferon responses and antiviral immunity. Cell Mol. Immunol. 2018, 15, 48–57. [Google Scholar] [CrossRef] [PubMed]
  44. Vidotto, T.; Saggioro, F.P.; Jamaspishvili, T.; Chesca, D.L.; Picanco de Albuquerque, C.G.; Reis, R.B.; Graham, C.H.; Berman, D.M.; Siemens, D.R.; Squire, J.A.; et al. PTEN-deficient prostate cancer is associated with an immunosuppressive tumor microenvironment mediated by increased expression of IDO1 and infiltrating FoxP3+ T regulatory cells. Prostate 2019, 79, 969–979. [Google Scholar] [CrossRef] [PubMed]
  45. Calcinotto, A.; Spataro, C.; Zagato, E.; Di Mitri, D.; Gil, V.; Crespo, M.; De Bernardis, G.; Losa, M.; Mirenda, M.; Pasquini, E.; et al. IL-23 secreted by myeloid cells drives castration-resistant prostate cancer. Nature 2018, 559, 363–369. [Google Scholar] [CrossRef] [PubMed]
  46. Zhao, D.; Cai, L.; Lu, X.; Liang, X.; Li, J.; Chen, P.; Ittmann, M.; Shang, X.; Jiang, S.; Li, H.; et al. Chromatin Regulator CHD1 Remodels the Immunosuppressive Tumor Microenvironment in PTEN-Deficient Prostate Cancer. Cancer Discov. 2020, 10, 1374–1387. [Google Scholar] [CrossRef] [PubMed]
  47. Formaggio, N.; Rubin, M.A.; Theurillat, J.P. Loss and revival of androgen receptor signaling in advanced prostate cancer. Oncogene 2021, 40, 1205–1216. [Google Scholar] [CrossRef] [PubMed]
  48. Crawford, E.D.; Heidenreich, A.; Lawrentschuk, N.; Tombal, B.; Pompeo, A.C.L.; Mendoza-Valdes, A.; Miller, K.; Debruyne, F.M.J.; Klotz, L. Androgen-targeted therapy in men with prostate cancer: Evolving practice and future considerations. Prostate Cancer Prostatic. Dis. 2019, 22, 24–38. [Google Scholar] [CrossRef] [Green Version]
  49. Ben-Batalla, I.; Vargas-Delgado, M.E.; von Amsberg, G.; Janning, M.; Loges, S. Influence of Androgens on Immunity to Self and Foreign: Effects on Immunity and Cancer. Front. Immunol. 2020, 11, 1184. [Google Scholar] [CrossRef]
  50. Gamat, M.; McNeel, D.G. Androgen deprivation and immunotherapy for the treatment of prostate cancer. Endocr. Relat. Cancer 2017, 24, T297–T310. [Google Scholar] [CrossRef]
  51. Page, S.T.; Plymate, S.R.; Bremner, W.J.; Matsumoto, A.M.; Hess, D.L.; Lin, D.W.; Amory, J.K.; Nelson, P.S.; Wu, J.D. Effect of medical castration on CD4+ CD25+ T cells, CD8+ T cell IFN-gamma expression, and NK cells: A physiological role for testosterone and/or its metabolites. Am. J. Physiol. Endocrinol. Metab. 2006, 290, E856–E863. [Google Scholar] [CrossRef] [Green Version]
  52. Shen, Y.C.; Ghasemzadeh, A.; Kochel, C.M.; Nirschl, T.R.; Francica, B.J.; Lopez-Bujanda, Z.A.; Carrera Haro, M.A.; Tam, A.; Anders, R.A.; Selby, M.J.; et al. Combining intratumoral Treg depletion with androgen deprivation therapy (ADT): Preclinical activity in the Myc-CaP model. Prostate Cancer Prostatic. Dis. 2018, 21, 113–125. [Google Scholar] [CrossRef]
  53. Pu, Y.; Xu, M.; Liang, Y.; Yang, K.; Guo, Y.; Yang, X.; Fu, Y.X. Androgen receptor antagonists compromise T cell response against prostate cancer leading to early tumor relapse. Sci. Transl. Med. 2016, 8, 333ra347. [Google Scholar] [CrossRef] [Green Version]
  54. Obradovic, A.Z.; Dallos, M.C.; Zahurak, M.L.; Partin, A.W.; Schaeffer, E.M.; Ross, A.E.; Allaf, M.E.; Nirschl, T.R.; Liu, D.; Chapman, C.G.; et al. T-Cell Infiltration and Adaptive Treg Resistance in Response to Androgen Deprivation with or without Vaccination in Localized Prostate Cancer. Clin. Cancer Res. 2020, 26, 3182–3192. [Google Scholar] [CrossRef] [Green Version]
  55. Long, X.; Hou, H.; Wang, X.; Liu, S.; Diao, T.; Lai, S.; Hu, M.; Zhang, S.; Liu, M.; Zhang, H. Immune signature driven by ADT-induced immune microenvironment remodeling in prostate cancer is correlated with recurrence-free survival and immune infiltration. Cell Death Dis. 2020, 11, 779. [Google Scholar] [CrossRef]
  56. Shiao, S.L.; Chu, G.C.; Chung, L.W. Regulation of prostate cancer progression by the tumor microenvironment. Cancer Lett. 2016, 380, 340–348. [Google Scholar] [CrossRef] [Green Version]
  57. Stultz, J.; Fong, L. How to turn up the heat on the cold immune microenvironment of metastatic prostate cancer. Prostate Cancer Prostatic. Dis. 2021, 24, 697–717. [Google Scholar] [CrossRef]
  58. Chulpanova, D.S.; Kitaeva, K.V.; Green, A.R.; Rizvanov, A.A.; Solovyeva, V.V. Molecular Aspects and Future Perspectives of Cytokine-Based Anti-cancer Immunotherapy. Front. Cell Dev. Biol. 2020, 8, 402. [Google Scholar] [CrossRef]
  59. Mao, C.; Ding, Y.; Xu, N. A Double-Edged Sword Role of Cytokines in Prostate Cancer Immunotherapy. Front. Oncol. 2021, 11, 688489. [Google Scholar] [CrossRef]
  60. Wang, X.; Yang, L.; Huang, F.; Zhang, Q.; Liu, S.; Ma, L.; You, Z. Inflammatory cytokines IL-17 and TNF-α up-regulate PD-L1 expression in human prostate and colon cancer cells. Immunol. Lett. 2017, 184, 7–14. [Google Scholar] [CrossRef] [Green Version]
  61. Lundholm, M.; Hägglöf, C.; Wikberg, M.L.; Stattin, P.; Egevad, L.; Bergh, A.; Wikström, P.; Palmqvist, R.; Edin, S. Secreted Factors from Colorectal and Prostate Cancer Cells Skew the Immune Response in Opposite Directions. Sci. Rep. 2015, 5, 15651. [Google Scholar] [CrossRef]
  62. Wise, G.J.; Marella, V.K.; Talluri, G.; Shirazian, D. Cytokine variations in patients with hormone treated prostate cancer. J. Urol. 2000, 164, 722–725. [Google Scholar] [CrossRef]
  63. Mariathasan, S.; Turley, S.J.; Nickles, D.; Castiglioni, A.; Yuen, K.; Wang, Y.; Kadel Iii, E.E.; Koeppen, H.; Astarita, J.L.; Cubas, R.; et al. TGFβ attenuates tumour response to PD-L1 blockade by contributing to exclusion of T cells. Nature 2018, 554, 544–548. [Google Scholar] [CrossRef]
  64. Kloss, C.C.; Lee, J.; Zhang, A.; Chen, F.; Melenhorst, J.J.; Lacey, S.F.; Maus, M.V.; Fraietta, J.A.; Zhao, Y.; June, C.H. Dominant-Negative TGF-β Receptor Enhances PSMA-Targeted Human CAR T Cell Proliferation And Augments Prostate Cancer Eradication. Mol. Ther. 2018, 26, 1855–1866. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Tesi, R.J. MDSC; the Most Important Cell You Have Never Heard Of. Trends Pharm. Sci. 2019, 40, 4–7. [Google Scholar] [CrossRef]
  66. Koinis, F.; Xagara, A.; Chantzara, E.; Leontopoulou, V.; Aidarinis, C.; Kotsakis, A. Myeloid-Derived Suppressor Cells in Prostate Cancer: Present Knowledge and Future Perspectives. Cells 2022, 11, 20. [Google Scholar] [CrossRef] [PubMed]
  67. De Cicco, P.; Ercolano, G.; Ianaro, A. The New Era of Cancer Immunotherapy: Targeting Myeloid-Derived Suppressor Cells to Overcome Immune Evasion. Front. Immunol. 2020, 11, 1680. [Google Scholar] [CrossRef] [PubMed]
  68. Kwon, J.T.W.; Bryant, R.J.; Parkes, E.E. The tumor microenvironment and immune responses in prostate cancer patients. Endocr. Relat. Cancer 2021, 28, T95–T107. [Google Scholar] [CrossRef]
  69. Bezzi, M.; Seitzer, N.; Ishikawa, T.; Reschke, M.; Chen, M.; Wang, G.; Mitchell, C.; Ng, C.; Katon, J.; Lunardi, A.; et al. Diverse genetic-driven immune landscapes dictate tumor progression through distinct mechanisms. Nat. Med. 2018, 24, 165–175. [Google Scholar] [CrossRef]
  70. Hossain, D.M.; Pal, S.K.; Moreira, D.; Duttagupta, P.; Zhang, Q.; Won, H.; Jones, J.; D’Apuzzo, M.; Forman, S.; Kortylewski, M. TLR9-Targeted STAT3 Silencing Abrogates Immunosuppressive Activity of Myeloid-Derived Suppressor Cells from Prostate Cancer Patients. Clin. Cancer Res. 2015, 21, 3771–3782. [Google Scholar] [CrossRef] [Green Version]
  71. Idorn, M.; Køllgaard, T.; Kongsted, P.; Sengeløv, L.; Thor Straten, P. Correlation between frequencies of blood monocytic myeloid-derived suppressor cells, regulatory T cells and negative prognostic markers in patients with castration-resistant metastatic prostate cancer. Cancer Immunol. Immunother. 2014, 63, 1177–1187. [Google Scholar] [CrossRef]
  72. Lopez-Bujanda, Z.; Drake, C.G. Myeloid-derived cells in prostate cancer progression: Phenotype and prospective therapies. J. Leukoc. Biol 2017, 102, 393–406. [Google Scholar] [CrossRef] [Green Version]
  73. Wen, J.; Huang, G.; Liu, S.; Wan, J.; Wang, X.; Zhu, Y.; Kaliney, W.; Zhang, C.; Cheng, L.; Wen, X.; et al. Polymorphonuclear MDSCs are enriched in the stroma and expanded in metastases of prostate cancer. J. Pathol. Clin. Res. 2020, 6, 171–177. [Google Scholar] [CrossRef]
  74. Lu, X.; Horner, J.W.; Paul, E.; Shang, X.; Troncoso, P.; Deng, P.; Jiang, S.; Chang, Q.; Spring, D.J.; Sharma, P.; et al. Effective combinatorial immunotherapy for castration-resistant prostate cancer. Nature 2017, 543, 728–732. [Google Scholar] [CrossRef] [Green Version]
  75. Sternberg, C.; Armstrong, A.; Pili, R.; Ng, S.; Huddart, R.; Agarwal, N.; Khvorostenko, D.; Lyulko, O.; Brize, A.; Vogelzang, N.; et al. Randomized, Double-Blind, Placebo-Controlled Phase III Study of Tasquinimod in Men with Metastatic Castration-Resistant Prostate Cancer. J. Clin. Oncol. 2016, 34, 2636–2643. [Google Scholar] [CrossRef] [PubMed]
  76. Larionova, I.; Tuguzbaeva, G.; Ponomaryova, A.; Stakheyeva, M.; Cherdyntseva, N.; Pavlov, V.; Choinzonov, E.; Kzhyshkowska, J. Tumor-Associated Macrophages in Human Breast, Colorectal, Lung, Ovarian and Prostate Cancers. Front. Oncol. 2020, 10, 566511. [Google Scholar] [CrossRef]
  77. Erlandsson, A.; Carlsson, J.; Lundholm, M.; Fält, A.; Andersson, S.-O.; Andrén, O.; Davidsson, S. M2 macrophages and regulatory T cells in lethal prostate cancer. Prostate 2019, 79, 363–369. [Google Scholar] [CrossRef]
  78. Mantovani, A.; Marchesi, F.; Malesci, A.; Laghi, L.; Allavena, P. Tumour-associated macrophages as treatment targets in oncology. Nat. Rev. Clin. Oncol. 2017, 14, 399–416. [Google Scholar] [CrossRef]
  79. Lissbrant, I.F.; Stattin, P.; Wikstrom, P.; Damber, J.E.; Egevad, L.; Bergh, A. Tumor associated macrophages in human prostate cancer: Relation to clinicopathological variables and survival. Int. J. Oncol. 2000, 17, 445–496. [Google Scholar] [CrossRef]
  80. Shimura, S.; Yang, G.; Ebara, S.; Wheeler, T.M.; Frolov, A.; Thompson, T.C. Reduced Infiltration of Tumor-associated Macrophages in Human Prostate Cancer: Association with Cancer Progression1. Cancer Res. 2000, 60, 5857–5861. [Google Scholar]
  81. Chen, S.; Zhu, G.; Yang, Y.; Wang, F.; Xiao, Y.-T.; Zhang, N.; Bian, X.; Zhu, Y.; Yu, Y.; Liu, F.; et al. Single-cell analysis reveals transcriptomic remodellings in distinct cell types that contribute to human prostate cancer progression. Nat. Cell Biol. 2021, 23, 87–98. [Google Scholar] [CrossRef] [PubMed]
  82. Thienger, P.; Rubin, M.A. Prostate cancer hijacks the microenvironment. Nat. Cell Biol. 2021, 23, 3–5. [Google Scholar] [CrossRef] [PubMed]
  83. Seif, F.; Sharifi, L.; Khoshmirsafa, M.; Mojibi, Y.; Mohsenzadegan, M. A Review of Preclinical Experiments toward Targeting M2 Macrophages in Prostate Cancer. Curr. Drug Targets 2019, 20, 789–798. [Google Scholar] [CrossRef] [PubMed]
  84. Joyce, J.A.; Pollard, J.W. Microenvironmental regulation of metastasis. Nat. Rev. Cancer 2009, 9, 239–252. [Google Scholar] [CrossRef]
  85. Deryugina, E.I.; Kiosses, W.B. Intratumoral Cancer Cell Intravasation Can Occur Independent of Invasion into the Adjacent Stroma. Cell Rep. 2017, 19, 601–616. [Google Scholar] [CrossRef] [Green Version]
  86. Bonollo, F.; Thalmann, G.N.; Kruithof-de Julio, M.; Karkampouna, S. The Role of Cancer-Associated Fibroblasts in Prostate Cancer Tumorigenesis. Cancers 2020, 12, 1887. [Google Scholar] [CrossRef]
  87. Monteran, L.; Erez, N. The Dark Side of Fibroblasts: Cancer-Associated Fibroblasts as Mediators of Immunosuppression in the Tumor Microenvironment. Front. Immunol. 2019, 10, 1835. [Google Scholar] [CrossRef] [Green Version]
  88. Giannoni, E.; Bianchini, F.; Masieri, L.; Serni, S.; Torre, E.; Calorini, L.; Chiarugi, P. Reciprocal activation of prostate cancer cells and cancer-associated fibroblasts stimulates epithelial-mesenchymal transition and cancer stemness. Cancer Res. 2010, 70, 6945–6956. [Google Scholar] [CrossRef] [Green Version]
  89. Fukumura, D.; Kloepper, J.; Amoozgar, Z.; Duda, D.G.; Jain, R.K. Enhancing cancer immunotherapy using antiangiogenics: Opportunities and challenges. Nat. Rev. Clin. Oncol. 2018, 15, 325–340. [Google Scholar] [CrossRef]
  90. Ziani, L.; Safta-Saadoun, T.B.; Gourbeix, J.; Cavalcanti, A.; Robert, C.; Favre, G.; Chouaib, S.; Thiery, J. Melanoma-associated fibroblasts decrease tumor cell susceptibility to NK cell-mediated killing through matrix-metalloproteinases secretion. Oncotarget 2017, 8, 19780–19794. [Google Scholar] [CrossRef] [Green Version]
  91. Li, T.; Yang, Y.; Hua, X.; Wang, G.; Liu, W.; Jia, C.; Tai, Y.; Zhang, Q.; Chen, G. Hepatocellular carcinoma-associated fibroblasts trigger NK cell dysfunction via PGE2 and IDO. Cancer Lett. 2012, 318, 154–161. [Google Scholar] [CrossRef]
  92. Ohta, A.; Sitkovsky, M. Role of G-protein-coupled adenosine receptors in downregulation of inflammation and protection from tissue damage. Nature 2001, 414, 916–920. [Google Scholar] [CrossRef] [Green Version]
  93. Beavis, P.A.; Stagg, J.; Darcy, P.K.; Smyth, M.J. CD73: A potent suppressor of antitumor immune responses. Trends Immunol. 2012, 33, 231–237. [Google Scholar] [CrossRef]
  94. Stagg, J.; Divisekera, U.; Duret, H.; Sparwasser, T.; Teng, M.W.; Darcy, P.K.; Smyth, M.J. CD73-deficient mice have increased antitumor immunity and are resistant to experimental metastasis. Cancer Res. 2011, 71, 2892–2900. [Google Scholar] [CrossRef] [Green Version]
  95. Li, J.; Wang, L.; Chen, X.; Li, L.; Li, Y.; Ping, Y.; Huang, L.; Yue, D.; Zhang, Z.; Wang, F.; et al. CD39/CD73 upregulation on myeloid-derived suppressor cells via TGF-β-mTOR-HIF-1 signaling in patients with non-small cell lung cancer. Oncoimmunology 2017, 6, e1320011. [Google Scholar] [CrossRef] [Green Version]
  96. Yu, M.; Guo, G.; Huang, L.; Deng, L.; Chang, C.S.; Achyut, B.R.; Canning, M.; Xu, N.; Arbab, A.S.; Bollag, R.J.; et al. CD73 on cancer-associated fibroblasts enhanced by the A(2B)-mediated feedforward circuit enforces an immune checkpoint. Nat. Commun. 2020, 11, 515. [Google Scholar] [CrossRef]
  97. Vigano, S.; Alatzoglou, D.; Irving, M.; Ménétrier-Caux, C.; Caux, C.; Romero, P.; Coukos, G. Targeting Adenosine in Cancer Immunotherapy to Enhance T-Cell Function. Front. Immunol. 2019, 10, 925. [Google Scholar] [CrossRef] [Green Version]
  98. Leone, R.D.; Emens, L.A. Targeting adenosine for cancer immunotherapy. J. Immunother. Cancer 2018, 6, 57. [Google Scholar] [CrossRef] [Green Version]
  99. Allard, B.; Longhi, M.S.; Robson, S.C.; Stagg, J. The ectonucleotidases CD39 and CD73: Novel checkpoint inhibitor targets. Immunol. Rev. 2017, 276, 121–144. [Google Scholar] [CrossRef] [Green Version]
  100. Graddis, T.J.; McMahan, C.J.; Tamman, J.; Page, K.J.; Trager, J.B. Prostatic acid phosphatase expression in human tissues. Int. J. Clin. Exp. Pathol. 2011, 4, 295–306. [Google Scholar]
  101. Bendell, J.; Bauer, T.; Patel, M.; Falchook, G.; Karlix, J.L.; Lim, E.; Mugundu, G.; Mitchell, P.D.; Pouliot, G.P.; Moorthy, G.; et al. Abstract CT026: Evidence of immune activation in the first-in-human Phase Ia dose escalation study of the adenosine 2a receptor antagonist, AZD4635, in patients with advanced solid tumors. Cancer Res. 2019, 79, CT026. [Google Scholar] [CrossRef]
  102. Wise, D.R.; Gardner, O.; Gilbert, H.N.; Rieger, A.; Paoloni, M.C.; Krishnan, K. A phase Ib/II, open-label, platform study evaluating the efficacy and safety of AB928-based treatment combinations in participants with metastatic castrate-resistant prostate cancer. J. Clin. Oncol. 2020, 38, TPS272. [Google Scholar] [CrossRef]
  103. Cha, H.R.; Lee, J.H.; Ponnazhagan, S. Revisiting Immunotherapy: A Focus on Prostate Cancer. Cancer Res. 2020, 80, 1615–1623. [Google Scholar] [CrossRef] [Green Version]
  104. Powers, E.; Karachaliou, G.S.; Kao, C.; Harrison, M.R.; Hoimes, C.J.; George, D.J.; Armstrong, A.J.; Zhang, T. Novel therapies are changing treatment paradigms in metastatic prostate cancer. J. Hematol. Oncol. 2020, 13, 144. [Google Scholar] [CrossRef] [PubMed]
  105. Maiorano, B.A.; Schinzari, G.; Ciardiello, D.; Rodriquenz, M.G.; Cisternino, A.; Tortora, G.; Maiello, E. Cancer Vaccines for Genitourinary Tumors: Recent Progresses and Future Possibilities. Vaccines 2021, 9, 623. [Google Scholar] [CrossRef] [PubMed]
  106. Arlen, P.M.; Mohebtash, M.; Madan, R.A.; Gulley, J.L. Promising novel immunotherapies and combinations for prostate cancer. Future Oncol. 2009, 5, 187–196. [Google Scholar] [CrossRef] [Green Version]
  107. Marshall, C.H.; Park, J.C.; DeWeese, T.L.; King, S.; Afful, M.; Hurrelbrink, J.; Manogue, C.; Cotogno, P.; Moldawer, N.P.; Barata, P.C.; et al. Randomized phase II study of sipuleucel-T (SipT) with or without radium-223 (Ra223) in men with asymptomatic bone-metastatic castrate-resistant prostate cancer (mCRPC). J. Clin. Oncol. 2020, 38, 130. [Google Scholar] [CrossRef]
  108. Sutherland, S.I.M.; Ju, X.; Horvath, L.G.; Clark, G.J. Moving on From Sipuleucel-T: New Dendritic Cell Vaccine Strategies for Prostate Cancer. Front. Immunol. 2021, 12, 641307. [Google Scholar] [CrossRef]
  109. Vogelzang, N.J.; Beer, T.M.; Gerritsen, W.; Oudard, S.; Wiechno, P.; Kukielka-Budny, B.; Samal, V.; Hajek, J.; Feyerabend, S.; Khoo, V.; et al. Efficacy and Safety of Autologous Dendritic Cell-Based Immunotherapy, Docetaxel, and Prednisone vs Placebo in Patients with Metastatic Castration-Resistant Prostate Cancer: The VIABLE Phase 3 Randomized Clinical Trial. JAMA Oncol. 2022, e217298, in press. [Google Scholar] [CrossRef]
  110. Schuhmacher, J.; Heidu, S.; Balchen, T.; Richardson, J.R.; Schmeltz, C.; Sonne, J.; Schweiker, J.; Rammensee, H.G.; Thor Straten, P.; Røder, M.A.; et al. Vaccination against RhoC induces long-lasting immune responses in patients with prostate cancer: Results from a phase I/II clinical trial. J. Immunother Cancer 2020, 8, e001157. [Google Scholar] [CrossRef]
  111. Lilleby, W.; Gaudernack, G.; Brunsvig, P.F.; Vlatkovic, L.; Schulz, M.; Mills, K.; Hole, K.H.; Inderberg, E.M. Phase I/IIa clinical trial of a novel hTERT peptide vaccine in men with metastatic hormone-naive prostate cancer. Cancer Immunol. Immunother. 2017, 66, 891–901. [Google Scholar] [CrossRef]
  112. Yoshimura, K.; Minami, T.; Nozawa, M.; Kimura, T.; Egawa, S.; Fujimoto, H.; Yamada, A.; Itoh, K.; Uemura, H. A Phase 2 Randomized Controlled Trial of Personalized Peptide Vaccine Immunotherapy with Low-dose Dexamethasone Versus Dexamethasone Alone in Chemotherapy-naive Castration-resistant Prostate Cancer. Eur. Urol. 2016, 70, 35–41. [Google Scholar] [CrossRef] [Green Version]
  113. Kantoff, P.W.; Schuetz, T.J.; Blumenstein, B.A.; Glode, L.M.; Bilhartz, D.L.; Wyand, M.; Manson, K.; Panicali, D.L.; Laus, R.; Schlom, J.; et al. Overall survival analysis of a phase II randomized controlled trial of a Poxviral-based PSA-targeted immunotherapy in metastatic castration-resistant prostate cancer. J. Clin. Oncol. 2010, 28, 1099–1105. [Google Scholar] [CrossRef]
  114. Gulley, J.L.; Borre, M.; Vogelzang, N.J.; Ng, S.; Agarwal, N.; Parker, C.C.; Pook, D.W.; Rathenborg, P.; Flaig, T.W.; Carles, J.; et al. Phase III Trial of PROSTVAC in Asymptomatic or Minimally Symptomatic Metastatic Castration-Resistant Prostate Cancer. J. Clin. Oncol. 2019, 37, 1051–1061. [Google Scholar] [CrossRef]
  115. Zhou, S.; Gravekamp, C.; Bermudes, D.; Liu, K. Tumour-targeting bacteria engineered to fight cancer. Nat. Rev. Cancer 2018, 18, 727–743. [Google Scholar] [CrossRef]
  116. Gupta, K.H.; Nowicki, C.; Giurini, E.F.; Marzo, A.L.; Zloza, A. Bacterial-Based Cancer Therapy (BBCT): Recent Advances, Current Challenges, and Future Prospects for Cancer Immunotherapy. Vaccines 2021, 9, 1497. [Google Scholar] [CrossRef]
  117. Stein, M.N.; Fong, L.; Mega, A.E.; Lam, E.T.; Heyburn, J.W.; GUTIERREZ, A.A.; Parsi, M.; Vangala, S.; Haas, N.B. KEYNOTE-046 (Part B): Effects of ADXS-PSA in combination with pembrolizumab on survival in metastatic, castration-resistant prostate cancer patients with or without prior exposure to docetaxel. J. Clin. Oncol. 2020, 38, 126. [Google Scholar] [CrossRef]
  118. Gamat-Huber, M.; Jeon, D.; Johnson, L.E.; Moseman, J.E.; Muralidhar, A.; Potluri, H.K.; Rastogi, I.; Wargowski, E.; Zahm, C.D.; McNeel, D.G. Treatment Combinations with DNA Vaccines for the Treatment of Metastatic Castration-Resistant Prostate Cancer (mCRPC). Cancers 2020, 12, 2831. [Google Scholar] [CrossRef]
  119. McNeel, D.G.; Eickhoff, J.C.; Johnson, L.E.; Roth, A.R.; Perk, T.G.; Fong, L.; Antonarakis, E.S.; Wargowski, E.; Jeraj, R.; Liu, G. Phase II Trial of a DNA Vaccine Encoding Prostatic Acid Phosphatase (pTVG-HP [MVI-816]) in Patients with Progressive, Nonmetastatic, Castration-Sensitive Prostate Cancer. J. Clin. Oncol. 2019, 37, 3507–3517. [Google Scholar] [CrossRef]
  120. Shore, N.D.; Morrow, M.P.; McMullan, T.; Kraynyak, K.A.; Sylvester, A.; Bhatt, K.; Cheung, J.; Boyer, J.D.; Liu, L.; Sacchetta, B.; et al. CD8(+) T Cells Impact Rising PSA in Biochemically Relapsed Cancer Patients Using Immunotherapy Targeting Tumor-Associated Antigens. Mol. Ther. 2020, 28, 1238–1250. [Google Scholar] [CrossRef]
  121. Kyriakopoulos, C.E.; Eickhoff, J.C.; Ferrari, A.C.; Schweizer, M.T.; Wargowski, E.; Olson, B.M.; McNeel, D.G. Multicenter Phase I Trial of a DNA Vaccine Encoding the Androgen Receptor Ligand-binding Domain (pTVG-AR, MVI-118) in Patients with Metastatic Prostate Cancer. Clin. Cancer Res. 2020, 26, 5162–5171. [Google Scholar] [CrossRef] [PubMed]
  122. Stenzl, A.; Feyerabend, S.; Syndikus, I.; Sarosiek, T.; Kübler, H.; Heidenreich, A.; Cathomas, R.; Grüllich, C.; Loriot, Y.; Perez Gracia, S.L.; et al. Results of the randomized, placebo-controlled phase I/IIB trial of CV9104, an mRNA based cancer immunotherapy, in patients with metastatic castration-resistant prostate cancer (mCRPC). Ann. Oncol. 2017, 28, v408–v409. [Google Scholar] [CrossRef]
  123. Beer, T.M.; Kwon, E.D.; Drake, C.G.; Fizazi, K.; Logothetis, C.; Gravis, G.; Ganju, V.; Polikoff, J.; Saad, F.; Humanski, P.; et al. Randomized, Double-Blind, Phase III Trial of Ipilimumab Versus Placebo in Asymptomatic or Minimally Symptomatic Patients with Metastatic Chemotherapy-Naive Castration-Resistant Prostate Cancer. J. Clin. Oncol. 2017, 35, 40–47. [Google Scholar] [CrossRef] [PubMed]
  124. Kwon, E.D.; Drake, C.G.; Scher, H.I.; Fizazi, K.; Bossi, A.; van den Eertwegh, A.J.; Krainer, M.; Houede, N.; Santos, R.; Mahammedi, H.; et al. Ipilimumab versus placebo after radiotherapy in patients with metastatic castration-resistant prostate cancer that had progressed after docetaxel chemotherapy (CA184-043): A multicentre, randomised, double-blind, phase 3 trial. Lancet Oncol. 2014, 15, 700–712. [Google Scholar] [CrossRef] [Green Version]
  125. Cabel, L.; Loir, E.; Gravis, G.; Lavaud, P.; Massard, C.; Albiges, L.; Baciarello, G.; Loriot, Y.; Fizazi, K. Long-term complete remission with Ipilimumab in metastatic castrate-resistant prostate cancer: Case report of two patients. J. Immunother. Cancer 2017, 5, 31. [Google Scholar] [CrossRef] [Green Version]
  126. Basnet, A.; Khullar, G.; Mehta, R.; Chittoria, N. A Case of Locally Advanced Castration-resistant Prostate Cancer with Remarkable Response to Nivolumab. Clin. Genitourin. Cancer 2017, 15, e881–e884. [Google Scholar] [CrossRef]
  127. Fizazi, K.; Drake, C.G.; Beer, T.M.; Kwon, E.D.; Scher, H.I.; Gerritsen, W.R.; Bossi, A.; den Eertwegh, A.; Krainer, M.; Houede, N.; et al. Final Analysis of the Ipilimumab Versus Placebo Following Radiotherapy Phase III Trial in Postdocetaxel Metastatic Castration-resistant Prostate Cancer Identifies an Excess of Long-term Survivors. Eur. Urol. 2020, 78, 822–830. [Google Scholar] [CrossRef]
  128. Subudhi, S.K.; Vence, L.; Zhao, H.; Blando, J.; Yadav, S.S.; Xiong, Q.; Reuben, A.; Aparicio, A.; Corn, P.G.; Chapin, B.F.; et al. Neoantigen responses, immune correlates, and favorable outcomes after ipilimumab treatment of patients with prostate cancer. Sci. Transl. Med. 2020, 12, eaaz3577. [Google Scholar] [CrossRef]
  129. Antonarakis, E.S.; Piulats, J.M.; Gross-Goupil, M.; Goh, J.; Ojamaa, K.; Hoimes, C.J.; Vaishampayan, U.; Berger, R.; Sezer, A.; Alanko, T.; et al. Pembrolizumab for Treatment-Refractory Metastatic Castration-Resistant Prostate Cancer: Multicohort, Open-Label Phase II KEYNOTE-199 Study. J. Clin. Oncol. 2020, 38, 395–405. [Google Scholar] [CrossRef]
  130. Topalian, S.L.; Hodi, F.S.; Brahmer, J.R.; Gettinger, S.N.; Smith, D.C.; McDermott, D.F.; Powderly, J.D.; Carvajal, R.D.; Sosman, J.A.; Atkins, M.B.; et al. Safety, activity, and immune correlates of anti-PD-1 antibody in cancer. N. Engl. J. Med. 2012, 366, 2443–2454. [Google Scholar] [CrossRef]
  131. Hansen, A.R.; Massard, C.; Ott, P.A.; Haas, N.B.; Lopez, J.S.; Ejadi, S.; Wallmark, J.M.; Keam, B.; Delord, J.P.; Aggarwal, R.; et al. Pembrolizumab for advanced prostate adenocarcinoma: Findings of the KEYNOTE-028 study. Ann. Oncol. 2018, 29, 1807–1813. [Google Scholar] [CrossRef]
  132. Petrylak, D.P.; Loriot, Y.; Shaffer, D.R.; Braiteh, F.; Powderly, J.; Harshman, L.C.; Conkling, P.; Delord, J.P.; Gordon, M.; Kim, J.W.; et al. Safety and Clinical Activity of Atezolizumab in Patients with Metastatic Castration-Resistant Prostate Cancer: A Phase I Study. Clin. Cancer Res. 2021, 27, 3360–3369. [Google Scholar] [CrossRef]
  133. Antonarakis, E.S.; Isaacsson Velho, P.; Fu, W.; Wang, H.; Agarwal, N.; Sacristan Santos, V.; Maughan, B.L.; Pili, R.; Adra, N.; Sternberg, C.N.; et al. CDK12-Altered Prostate Cancer: Clinical Features and Therapeutic Outcomes to Standard Systemic Therapies, Poly (ADP-Ribose) Polymerase Inhibitors, and PD-1 Inhibitors. JCO Precis. Oncol. 2020, 4, 370–381. [Google Scholar] [CrossRef]
  134. Gao, J.; Ward, J.F.; Pettaway, C.A.; Shi, L.Z.; Subudhi, S.K.; Vence, L.M.; Zhao, H.; Chen, J.; Chen, H.; Efstathiou, E.; et al. VISTA is an inhibitory immune checkpoint that is increased after ipilimumab therapy in patients with prostate cancer. Nat. Med. 2017, 23, 551–555. [Google Scholar] [CrossRef] [Green Version]
  135. Subudhi, S.K.; Siddiqui, B.A.; Aparicio, A.M.; Yadav, S.S.; Basu, S.; Chen, H.; Jindal, S.; Tidwell, R.S.S.; Varma, A.; Logothetis, C.J.; et al. Combined CTLA-4 and PD-L1 blockade in patients with chemotherapy-naive metastatic castration-resistant prostate cancer is associated with increased myeloid and neutrophil immune subsets in the bone microenvironment. J. Immunother. Cancer 2021, 9, e002919. [Google Scholar] [CrossRef]
  136. Sharma, P.; Pachynski, R.K.; Narayan, V.; Flechon, A.; Gravis, G.; Galsky, M.D.; Mahammedi, H.; Patnaik, A.; Subudhi, S.K.; Ciprotti, M.; et al. Nivolumab Plus Ipilimumab for Metastatic Castration-Resistant Prostate Cancer: Preliminary Analysis of Patients in the CheckMate 650 Trial. Cancer Cell 2020, 38, 489–499.e483. [Google Scholar] [CrossRef]
  137. Boudadi, K.; Suzman, D.L.; Anagnostou, V.; Fu, W.; Luber, B.; Wang, H.; Niknafs, N.; White, J.R.; Silberstein, J.L.; Sullivan, R.; et al. Ipilimumab plus nivolumab and DNA-repair defects in AR-V7-expressing metastatic prostate cancer. Oncotarget 2018, 9, 28561–28571. [Google Scholar] [CrossRef] [Green Version]
  138. Shenderov, E.; Boudadi, K.; Fu, W.; Wang, H.; Sullivan, R.; Jordan, A.; Dowling, D.; Harb, R.; Schonhoft, J.; Jendrisak, A.; et al. Nivolumab plus ipilimumab, with or without enzalutamide, in AR-V7-expressing metastatic castration-resistant prostate cancer: A phase-2 nonrandomized clinical trial. Prostate 2021, 81, 326–338. [Google Scholar] [CrossRef]
  139. Graff, J.N.; Beer, T.M.; Alumkal, J.J.; Slottke, R.E.; Redmond, W.L.; Thomas, G.V.; Thompson, R.F.; Wood, M.A.; Koguchi, Y.; Chen, Y.; et al. A phase II single-arm study of pembrolizumab with enzalutamide in men with metastatic castration-resistant prostate cancer progressing on enzalutamide alone. J. Immunother. Cancer 2020, 8, e000642. [Google Scholar] [CrossRef]
  140. Hoimes, C.J.; Graff, J.N.; Tagawa, S.T.; Hwang, C.; Kilari, D.; Tije, A.J.T.; Omlin, A.; McDermott, R.S.; Vaishampayan, U.N.; Elliott, T.; et al. KEYNOTE-199 cohorts (C) 4 and 5: Phase II study of pembrolizumab (pembro) plus enzalutamide (enza) for enza-resistant metastatic castration-resistant prostate cancer (mCRPC). J. Clin. Oncol. 2020, 38, 5543. [Google Scholar] [CrossRef]
  141. Sweeney, C.J.; Gillessen, S.; Rathkopf, D.; Matsubara, N.; Drake, C.; Fizazi, K.; Piulats, J.M.; Wysocki, P.J.; Buchschacher, G.L.; Doss, J.; et al. Abstract CT014: IMbassador250: A phase III trial comparing atezolizumab with enzalutamide vs enzalutamide alone in patients with metastatic castration-resistant prostate cancer (mCRPC). Cancer Res. 2020, 80, CT014. [Google Scholar] [CrossRef]
  142. Galluzzi, L.; Buque, A.; Kepp, O.; Zitvogel, L.; Kroemer, G. Immunological Effects of Conventional Chemotherapy and Targeted Anticancer Agents. Cancer Cell 2015, 28, 690–714. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Fizazi, K.; González Mella, P.; Castellano, D.; Minatta, J.N.; Rezazadeh, A.; Shaffer, D.R.; Vazquez Limon, J.C.; Sánchez López, H.M.; Armstrong, A.J.; Horvath, L.; et al. CheckMate 9KD Arm B final analysis: Efficacy and safety of nivolumab plus docetaxel for chemotherapy-naïve metastatic castration-resistant prostate cancer. Rev. Clin. Oncol. 2021, 39, 12. [Google Scholar] [CrossRef]
  144. Fizazi, K.; Gonzalez Mella, P.; Castellano, D.; Minatta, J.N.; Rezazadeh Kalebasty, A.; Shaffer, D.; Vazquez Limon, J.C.; Sanchez Lopez, H.M.; Armstrong, A.J.; Horvath, L.; et al. Nivolumab plus docetaxel in patients with chemotherapy-naive metastatic castration-resistant prostate cancer: Results from the phase II CheckMate 9KD trial. Eur. J. Cancer 2022, 160, 61–71. [Google Scholar] [CrossRef]
  145. Strickland, K.C.; Howitt, B.E.; Shukla, S.A.; Rodig, S.; Ritterhouse, L.L.; Liu, J.F.; Garber, J.E.; Chowdhury, D.; Wu, C.J.; D’Andrea, A.D.; et al. Association and prognostic significance of BRCA1/2-mutation status with neoantigen load, number of tumor-infiltrating lymphocytes and expression of PD-1/PD-L1 in high grade serous ovarian cancer. Oncotarget 2016, 7, 13587–13598. [Google Scholar] [CrossRef] [Green Version]
  146. Karzai, F.; VanderWeele, D.; Madan, R.A.; Owens, H.; Cordes, L.M.; Hankin, A.; Couvillon, A.; Nichols, E.; Bilusic, M.; Beshiri, M.L.; et al. Activity of durvalumab plus olaparib in metastatic castration-resistant prostate cancer in men with and without DNA damage repair mutations. J. ImmunoTher. Cancer 2018, 6, 141. [Google Scholar] [CrossRef]
  147. Pachynski, R.K.; Retz, M.; Goh, J.C.; Burotto, M.; Gravis, G.; Castellano, D.; Flechon, A.; Zschaebitz, S.; Shaffer, D.R.; Limon, J.C.V.; et al. CheckMate 9KD cohort A1 final analysis: Nivolumab (NIVO) + rucaparib for post-chemotherapy (CT) metastatic castration-resistant prostate cancer (mCRPC). J. Clin. Oncol. 2021, 39, 5044. [Google Scholar] [CrossRef]
  148. Petrylak, D.P.; Perez-Gracia, J.L.; Lacombe, L.; Bastos, D.A.; Mahammedi, H.; Kwan, E.M.; Zschäbitz, S.; Armstrong, A.J.; Pachynski, R.K.; Goh, J.C.; et al. 579MO CheckMate 9KD cohort A2 final analysis: Nivolumab (NIVO) + rucaparib for chemotherapy (CT)-naïve metastatic castration-resistant prostate cancer (mCRPC). In Proceedings of the 2021 ESMO Annual Congress, Online, 16–21 September 2021. [Google Scholar]
  149. Yakes, F.M.; Chen, J.; Tan, J.; Yamaguchi, K.; Shi, Y.; Yu, P.; Qian, F.; Chu, F.; Bentzien, F.; Cancilla, B.; et al. Cabozantinib (XL184), a novel MET and VEGFR2 inhibitor, simultaneously suppresses metastasis, angiogenesis, and tumor growth. Mol. Cancer Ther. 2011, 10, 2298–2308. [Google Scholar] [CrossRef] [Green Version]
  150. Kwilas, A.R.; Ardiani, A.; Donahue, R.N.; Aftab, D.T.; Hodge, J.W. Dual effects of a targeted small-molecule inhibitor (cabozantinib) on immune-mediated killing of tumor cells and immune tumor microenvironment permissiveness when combined with a cancer vaccine. J. Transl. Med. 2014, 12, 294. [Google Scholar] [CrossRef] [Green Version]
  151. Tolaney, S.M.; Ziehr, D.R.; Guo, H.; Ng, M.R.; Barry, W.T.; Higgins, M.J.; Isakoff, S.J.; Brock, J.E.; Ivanova, E.V.; Paweletz, C.P.; et al. Phase II and Biomarker Study of Cabozantinib in Metastatic Triple-Negative Breast Cancer Patients. Oncologist 2017, 22, 25–32. [Google Scholar] [CrossRef] [Green Version]
  152. Agarwal, N.M.B.; Maughan, B.L.; Dorff, T.; Kelly, W.; Fang, B.; McKay, R.; Singh, P.; Pagliaro, L.; Dreicer, R.; Srinivas, S.; et al. Cabozantinib (C) in combination with atezolizumab (A) in patients (pts) with metastatic castration-resistant prostate cancer (mCRPC): Results of expanded cohort 6 of the COSMIC-021 study. Ann. Oncol. 2021, 32, S1283–S1346. [Google Scholar] [CrossRef]
  153. Postow, M.A.; Callahan, M.K.; Barker, C.A.; Yamada, Y.; Yuan, J.; Kitano, S.; Mu, Z.; Rasalan, T.; Adamow, M.; Ritter, E.; et al. Immunologic correlates of the abscopal effect in a patient with melanoma. N. Engl. J. Med. 2012, 366, 925–931. [Google Scholar] [CrossRef] [Green Version]
  154. Formenti, S.C.; Rudqvist, N.P.; Golden, E.; Cooper, B.; Wennerberg, E.; Lhuillier, C.; Vanpouille-Box, C.; Friedman, K.; Ferrari de Andrade, L.; Wucherpfennig, K.W.; et al. Radiotherapy induces responses of lung cancer to CTLA-4 blockade. Nat. Med. 2018, 24, 1845–1851. [Google Scholar] [CrossRef]
  155. Fong, L.; Morris, M.J.; Sartor, O.; Higano, C.S.; Pagliaro, L.; Alva, A.; Appleman, L.J.; Tan, W.; Vaishampayan, U.; Porcu, R.; et al. A Phase Ib Study of Atezolizumab with Radium-223 Dichloride in Men with Metastatic Castration-Resistant Prostate Cancer. Clin. Cancer Res. 2021, 27, 4746–4756. [Google Scholar] [CrossRef]
  156. Einsele, H.; Borghaei, H.; Orlowski, R.Z.; Subklewe, M.; Roboz, G.J.; Zugmaier, G.; Kufer, P.; Iskander, K.; Kantarjian, H.M. The BiTE (bispecific T-cell engager) platform: Development and future potential of a targeted immuno-oncology therapy across tumor types. Cancer 2020, 126, 3192–3201. [Google Scholar] [CrossRef]
  157. Offner, S.; Hofmeister, R.; Romaniuk, A.; Kufer, P.; Baeuerle, P.A. Induction of regular cytolytic T cell synapses by bispecific single-chain antibody constructs on MHC class I-negative tumor cells. Mol. Immunol. 2006, 43, 763–771. [Google Scholar] [CrossRef]
  158. Dang, K.; Castello, G.; Clarke, S.C.; Li, Y.; Balasubramani, A.; Boudreau, A.; Davison, L.; Harris, K.E.; Pham, D.; Sankaran, P.; et al. Attenuating CD3 affinity in a PSMAxCD3 bispecific antibody enables killing of prostate tumor cells with reduced cytokine release. J. Immunother. Cancer 2021, 9, e002488. [Google Scholar] [CrossRef]
  159. Deegen, P.; Thomas, O.; Nolan-Stevaux, O.; Li, S.; Wahl, J.; Bogner, P.; Aeffner, F.; Friedrich, M.; Liao, M.Z.; Matthes, K.; et al. The PSMA-targeting Half-life Extended BiTE Therapy AMG 160 has Potent Antitumor Activity in Preclinical Models of Metastatic Castration-resistant Prostate Cancer. Clin. Cancer Res. 2021, 27, 2928–2937. [Google Scholar] [CrossRef]
  160. Tran, B.H.L.; Dorff, T.; Rettig, M.; Lolkema, M.P.; Machiels, J.P.; Rottey, S.; Autio, K.; Greil, R.; Adra, N.; Lemech, C.; et al. #6090 Interim results from a phase 1 study of AMG 160, a half-life extended (HLE), PSMA-targeted, bispecific T-cell engager (BiTE®) immune therapy for metastatic castration-resistant prostate cancer (mCRPC). In Proceedings of the ESMO 2020, Online, 19–21 September 2020. [Google Scholar]
  161. Lund, M.E.; Howard, C.B.; Thurecht, K.J.; Campbell, D.H.; Mahler, S.M.; Walsh, B.J. A bispecific T cell engager targeting Glypican-1 redirects T cell cytolytic activity to kill prostate cancer cells. BMC Cancer 2020, 20, 1214. [Google Scholar] [CrossRef]
  162. Yamamoto, K.; Trad, A.; Baumgart, A.; Huske, L.; Lorenzen, I.; Chalaris, A.; Grotzinger, J.; Dechow, T.; Scheller, J.; Rose-John, S. A novel bispecific single-chain antibody for ADAM17 and CD3 induces T-cell-mediated lysis of prostate cancer cells. Biochem. J. 2012, 445, 135–144. [Google Scholar] [CrossRef] [Green Version]
  163. Giffin, M.J.; Cooke, K.; Lobenhofer, E.K.; Estrada, J.; Zhan, J.; Deegen, P.; Thomas, M.; Murawsky, C.M.; Werner, J.; Liu, S.; et al. AMG 757, a Half-Life Extended, DLL3-Targeted Bispecific T-Cell Engager, Shows High Potency and Sensitivity in Preclinical Models of Small-Cell Lung Cancer. Clin. Cancer Res. 2021, 27, 1526–1537. [Google Scholar] [CrossRef]
  164. Puca, L.; Gavyert, K.; Sailer, V.; Conteduca, V.; Dardenne, E.; Sigouros, M.; Isse, K.; Kearney, M.; Vosoughi, A.; Fernandez, L.; et al. Delta-like protein 3 expression and therapeutic targeting in neuroendocrine prostate cancer. Sci. Transl. Med. 2019, 11, eaav0891. [Google Scholar] [CrossRef]
  165. Goebeler, M.E.; Bargou, R.C. T cell-engaging therapies—BiTEs and beyond. Nat. Rev. Clin. Oncol. 2020, 17, 418–434. [Google Scholar] [CrossRef]
  166. Belmontes, B.; Sawant, D.V.; Zhong, W.; Tan, H.; Kaul, A.; Aeffner, F.; O’Brien, S.A.; Chun, M.; Noubade, R.; Eng, J.; et al. Immunotherapy combinations overcome resistance to bispecific T cell engager treatment in T cell-cold solid tumors. Sci. Transl. Med. 2021, 13, eabd1524. [Google Scholar] [CrossRef] [PubMed]
  167. Zorko, N.A.; Ryan, C.J. Novel immune engagers and cellular therapies for metastatic castration-resistant prostate cancer: Do we take a BiTe or ride BiKEs, TriKEs, and CARs? Prostate Cancer Prostatic Dis. 2021, 24, 986–996. [Google Scholar] [CrossRef] [PubMed]
  168. Wolf, P.; Alzubi, J.; Gratzke, C.; Cathomen, T. The potential of CAR T cell therapy for prostate cancer. Nat. Rev. Urol. 2021, 18, 556–571. [Google Scholar] [CrossRef] [PubMed]
  169. Maude, S.L.; Frey, N.; Shaw, P.A.; Aplenc, R.; Barrett, D.M.; Bunin, N.J.; Chew, A.; Gonzalez, V.E.; Zheng, Z.; Lacey, S.F.; et al. Chimeric antigen receptor T cells for sustained remissions in leukemia. N. Engl. J. Med. 2014, 371, 1507–1517. [Google Scholar] [CrossRef] [Green Version]
  170. Schuster, S.J.; Bishop, M.R.; Tam, C.S.; Waller, E.K.; Borchmann, P.; McGuirk, J.P.; Jäger, U.; Jaglowski, S.; Andreadis, C.; Westin, J.R.; et al. Tisagenlecleucel in Adult Relapsed or Refractory Diffuse Large B-Cell Lymphoma. N. Engl. J. Med. 2019, 380, 45–56. [Google Scholar] [CrossRef]
  171. Munshi, N.C.; Anderson, L.D., Jr.; Shah, N.; Madduri, D.; Berdeja, J.; Lonial, S.; Raje, N.; Lin, Y.; Siegel, D.; Oriol, A.; et al. Idecabtagene Vicleucel in Relapsed and Refractory Multiple Myeloma. N. Engl. J. Med. 2021, 384, 705–716. [Google Scholar] [CrossRef]
  172. Porter, D.L.; Hwang, W.T.; Frey, N.V.; Lacey, S.F.; Shaw, P.A.; Loren, A.W.; Bagg, A.; Marcucci, K.T.; Shen, A.; Gonzalez, V.; et al. Chimeric antigen receptor T cells persist and induce sustained remissions in relapsed refractory chronic lymphocytic leukemia. Sci. Transl. Med. 2015, 7, 303ra139. [Google Scholar] [CrossRef] [Green Version]
  173. Bostwick, D.G.; Pacelli, A.; Blute, M.; Roche, P.; Murphy, G.P. Prostate specific membrane antigen expression in prostatic intraepithelial neoplasia and adenocarcinoma: A study of 184 cases. Cancer 1998, 82, 2256–2261. [Google Scholar] [CrossRef]
  174. Troyer, J.K.; Beckett, M.L.; Wright, G.L. Detection and characterization of the prostate-specific membrane antigen (PSMA) in tissue extracts and body fluids. Int. J. Cancer 1995, 62, 552–558. [Google Scholar] [CrossRef]
  175. Ristau, B.T.; O’Keefe, D.S.; Bacich, D.J. The prostate-specific membrane antigen: Lessons and current clinical implications from 20 years of research. Urol. Oncol. 2014, 32, 272–279. [Google Scholar] [CrossRef] [Green Version]
  176. Reiter, R.E.; Gu, Z.; Watabe, T.; Thomas, G.; Szigeti, K.; Davis, E.; Wahl, M.; Nisitani, S.; Yamashiro, J.; Le Beau, M.M.; et al. Prostate stem cell antigen: A cell surface marker overexpressed in prostate cancer. Proc. Natl. Acad. Sci. USA 1998, 95, 1735–1740. [Google Scholar] [CrossRef] [Green Version]
  177. Gu, Z.; Thomas, G.; Yamashiro, J.; Shintaku, I.P.; Dorey, F.; Raitano, A.; Witte, O.N.; Said, J.W.; Loda, M.; Reiter, R.E. Prostate stem cell antigen (PSCA) expression increases with high gleason score, advanced stage and bone metastasis in prostate cancer. Oncogene 2000, 19, 1288–1296. [Google Scholar] [CrossRef] [Green Version]
  178. Lawrence, M.G.; Lai, J.; Clements, J.A. Kallikreins on steroids: Structure, function, and hormonal regulation of prostate-specific antigen and the extended kallikrein locus. Endocr. Rev. 2010, 31, 407–446. [Google Scholar] [CrossRef]
  179. Lindo, L.; Wilkinson, L.H.; Hay, K.A. Befriending the Hostile Tumor Microenvironment in CAR T-Cell Therapy. Front. Immunol. 2020, 11, 618387. [Google Scholar] [CrossRef]
  180. Newick, K.; O’Brien, S.; Moon, E.; Albelda, S.M. CAR T Cell Therapy for Solid Tumors. Annu. Rev. Med. 2017, 68, 139–152. [Google Scholar] [CrossRef]
  181. Junghans, R.P.; Ma, Q.; Rathore, R.; Gomes, E.M.; Bais, A.J.; Lo, A.S.; Abedi, M.; Davies, R.A.; Cabral, H.J.; Al-Homsi, A.S.; et al. 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, 1257–1270. [Google Scholar] [CrossRef]
  182. Zhang, Q.; Helfand, B.T.; Carneiro, B.A.; Qin, W.; Yang, X.J.; Lee, C.; Zhang, W.; Giles, F.J.; Cristofanilli, M.; Kuzel, T.M. Efficacy against Human Prostate Cancer by Prostate-specific Membrane Antigen-specific, Transforming Growth Factor-β Insensitive Genetically Targeted CD8(+) T-cells Derived from Patients with Metastatic Castrate-resistant Disease. Eur. Urol. 2018, 73, 648–652. [Google Scholar] [CrossRef]
  183. Carabasi, M.H.; McKean, M.; Stein, M.N.; Schweizer, M.T.; Luke, J.J.; Narayan, V.; Pachynski, R.K.; Parikh, R.A.; Zhang, J.; Fountaine, T.J.; et al. PSMA targeted armored chimeric antigen receptor (CAR) T-cells in patients with advanced mCRPC: A phase I experience. J. Clin. Oncol. 2021, 39, 2534. [Google Scholar] [CrossRef]
  184. Zheng, Y.; Nandakumar, K.S.; Cheng, K. Optimization of CAR-T Cell-Based Therapies Using Small-Molecule-Based Safety Switches. J. Med. Chem. 2021, 64, 9577–9591. [Google Scholar] [CrossRef]
  185. Casucci, M.; Falcone, L.; Camisa, B.; Norelli, M.; Porcellini, S.; Stornaiuolo, A.; Ciceri, F.; Traversari, C.; Bordignon, C.; Bonini, C.; et al. Extracellular NGFR Spacers Allow Efficient Tracking and Enrichment of Fully Functional CAR-T Cells Co-Expressing a Suicide Gene. Front. Immunol. 2018, 9, 507. [Google Scholar] [CrossRef] [Green Version]
  186. Straathof, K.C.; Pulè, M.A.; Yotnda, P.; Dotti, G.; Vanin, E.F.; Brenner, M.K.; Heslop, H.E.; Spencer, D.M.; Rooney, C.M. An inducible caspase 9 safety switch for T-cell therapy. Blood 2005, 105, 4247–4254. [Google Scholar] [CrossRef] [PubMed]
  187. Brudno, J.N.; Somerville, R.P.; Shi, V.; Rose, J.J.; Halverson, D.C.; Fowler, D.H.; Gea-Banacloche, J.C.; Pavletic, S.Z.; Hickstein, D.D.; Lu, T.L.; et al. Allogeneic T Cells That Express an Anti-CD19 Chimeric Antigen Receptor Induce Remissions of B-Cell Malignancies That Progress after Allogeneic Hematopoietic Stem-Cell Transplantation without Causing Graft-Versus-Host Disease. J. Clin. Oncol. 2016, 34, 1112–1121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Depil, S.; Duchateau, P.; Grupp, S.A.; Mufti, G.; Poirot, L. ’Off-the-shelf’ allogeneic CAR T cells: Development and challenges. Nat. Rev. Drug Discov. 2020, 19, 185–199. [Google Scholar] [CrossRef] [PubMed]
  189. Martínez Bedoya, D.; Dutoit, V.; Migliorini, D. Allogeneic CAR T Cells: An Alternative to Overcome Challenges of CAR T Cell Therapy in Glioblastoma. Front. Immunol. 2021, 12, 640082. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Immunosuppressive and -stimulating factors influencing immune response in advanced prostate cancer. Abbreviations: T-reg: regulatory T cell, MHC I: major histocompatibility complex I, ADT: androgen-deprivation therapy, MDSC: myeloid-derived suppressor cell, TAM: tumor-associated macrophage, CAF: cancer-associated fibroblast, PTEN: phosphatase and tensin homolog, NK cell: Natural Killer cell, MSI: microsatellite instability, CTL: cytotoxic T lymphocyte, TMB: tumor mutational burden, HRR: homologous recombination repair, and MMR: mismatch repair.
Figure 1. Immunosuppressive and -stimulating factors influencing immune response in advanced prostate cancer. Abbreviations: T-reg: regulatory T cell, MHC I: major histocompatibility complex I, ADT: androgen-deprivation therapy, MDSC: myeloid-derived suppressor cell, TAM: tumor-associated macrophage, CAF: cancer-associated fibroblast, PTEN: phosphatase and tensin homolog, NK cell: Natural Killer cell, MSI: microsatellite instability, CTL: cytotoxic T lymphocyte, TMB: tumor mutational burden, HRR: homologous recombination repair, and MMR: mismatch repair.
Ijms 23 02569 g001
Figure 2. Immunotherapeutic approaches in advanced prostate cancer.
Figure 2. Immunotherapeutic approaches in advanced prostate cancer.
Ijms 23 02569 g002
Table 1. Active trials examining vaccination strategies in advanced prostate cancer.
Table 1. Active trials examining vaccination strategies in advanced prostate cancer.
Trial NameTrial
Phase
Estimated Enrolment (pts)Experimental TherapyDisease StageRequired Pre-TreatmetPrimary EndpointNCT Number
Vaccination (Phase 1/2)
OVM-200-1001 FIH36OVM-200mCRPC or locally advancedAny first-line therapySafety, tolerabilityNCT05104515
UR1534 120Bcl-xl_42-CAF09bmHSCPADTSafetyNCT03412786
17-C-00071/229PROSTVAC-V/F + NivomCRPCADTSafetyNCT02933255
QuEST11/2113BN-Brachyury + M7824
vs.
BN-Brachyury + M7824 + N-803
vs.
BN-Brachyury + M7824 + N-803 + Epacadostat
mCRPC1 NHA or
if MSI high/MMRd: Pembrolizumab or if HRR mutation: Olaparib/Rucaparib
PSA decline of ≥30% (>21 days)
and/or
OR (RECIST 1.1)
NCT03493945
UW18037260pTVG-HP + Pembrolizumab
vs.
pTVG-HP + pTVG-AR + Pembrolizumab
mCRPCADTPFSNCT04090528
PRO-MERIT1/2130W_pro1
+ Cemiplimab
mCRPC2-3 linesDLTs
TEAEs
ORR (Part 2 Arms 1A and 1B)
NCT04382898
FIH, first in human; mCRPC, metastatic castration-resistant prostate cancer; mHNPC, metastatic hormone-naive prostate cancer; ADT, androgen-deprivation therapy; NHA, new hormonal agent; MSI, microsatellite instability; MMRd, mismatch repair deficiency; HRR, homologous recombination repair; DLT, dose-limiting toxicities; TEAEs, treatment-emergent adverse events.
Table 2. Active clinical trials on biomarker-selected patients and combinational treatment approaches with checkpoint inhibitors in advanced prostate cancer.
Table 2. Active clinical trials on biomarker-selected patients and combinational treatment approaches with checkpoint inhibitors in advanced prostate cancer.
Trial NameTrial PhaseEstimated
Enrolment
(pts)
Experimental TherapyDisease
Stage
Required PretreatmentPrimary EndpointNCT Number
CONTACT-023580Atezolizumab
Cabozantinib
mCRPC 1 NHA
docetaxel only in HNPC
Duration of PFS
by
RECIST1.1
NCT04446117
EVOLUTION2110Nivolumab + Ipilimumab + 177 Lu-PSMAmCRPCProgression on 1 NHAPSA-PFS at 1 yearNCT05150236
CheckMate7DX3984Nivolumab + Docetaxel followed by NivolumabmCRPC1-2 NHArPFS, OSNCT04100018
KEYNOTE-99131232Pembrolizumab +
Enzalutamide
mHNPCDocetaxel in HNPC allowedrPFS, OSNCT04191096
KEYNOTE-64131200Pembrolizumab +
Enzalutamide
mCRPCChemotherapy-naïve, abiraterone-naïve, or intolerant or progressed on abirateroneOS, rPFSNCT03834493
KEYLYNK-0103780Pembrolizumab + OlaparibmCRPC1 NHA and DocetaxelOS, rPFSNCT03834519
KEYNOTE-92131000Pembrolizumab +
Docetaxel
mCRPC≤1 NHA or mHSPC or mCRPCOS, rPFSNCT03834506
Trial NameTrial
Phase
Estimated
Enrolment
(pts)
Experimental TherapyDisease StageRequired PretreatmentPrimary Endpoint(s)NCT Number
AZD4635 in prostate cancer260Module 1: AZD4635 + durvalumab; Module 2: AZD4635 + oleclumabmCRPCProgressed on standard of careORR, PSA RR (>50%)NCT04089553
QUEST
(combination 1)
1b/2136Cetrelimab + NiraparibmCRPCnsPart 1: incidence of specific toxitities
Part 2: ORR
NCT03431350
KRONOS1b33Cetrelimab + ApalutamidmCRPCProgression on NHAAdverse events
PSA Response week 12
NCT03551782
ImmunoProst238Nivolumab+HRD mCRPC 1docetaxelPSA RR (>50%)NCT03040791
PORTER145A: Nivolumab + NKTR-214
B: Nivolumab + SBRT + CDX-301 + Poly-ICLC
C: Nivolumab + CDX-301 + INO-5151
mCRPCPrior NHA (e.g., abiraterone, enzalutamide, apalautamide)Incidence and severity of adverse eventsNCT03835533
201808043
CA209-9MW
120Nivolumab/Ipilimumab/
PROSTVAC/Neoantigen DNA vaccine
mHNPCChemohormonal therapySafety and TolerabilityNCT03532217
CA209-9352175Nivolumab + Ipilimumab (4 times) followed by Nivolumab maintenancemCRPC with immunogenic signature 21 line of therapyComposite response rate 3NCT03061539
IMPACT
CA209-8JJ
(cohort A)
240Nivolumab + Ipilimumab (4 times) followed by Nivolumab maintenancemCRPC with CDK12 mutationsnsPSA RR (>50%)NCT03570619
INSPIRE
CA184-585
275Nivolumab + Ipilimumab (4 times) followed by Nivolumab monotherapymCRPC with immunogenic phenotype 4nsDCR 5NCT04717154
Rad2Nivo
CA209-7G6
1b/236Nivolumab +
Radium223
Symptomatic
mCRPC without visceral Mets
nsSafety
ctDNA reduction after 6 weeks
NCT04109729
PLANE-PC250Pembrolizumab + LenvatinibNeuroendocrine PCansrPFSNCT04848337
Keynote 3651b/21000Cohort A AC: Pembrolizumab + Olaparib
Cohort B AC: Pembrolizumab + Docetaxel + Prednisone
Cohort C AC: Pembrolizumab + Enzalutamide
Cohort D AC: Pembrolizumab + Abiraterone + Prednisone
Cohort E AC: Pembrolizumab + Lenvatinib
Cohort F t-NE: Pembrolizumab + Lenvatinib
Cohort G (AC) Pembrolizumab/Vibostolimab coformulation
Cohort H t-NE: Pembrolizumab/Vibostolimab coformulation
Cohort I t-NE: Pembrolizumab + Carboplatin + Etoposide
For Cohorts A, B, C, D, E, and G: histologically or cytologically confirmed adenocarcinoma of the prostate without small cell histology
Cohorts F, H, and I: neuroendocrine PCa defined by ≥1% neuroendocrine cells in a recent biopsy specimen
Cohort E: Docetaxel + up to 2 NHA
Cohort F, G, H, I: Docetaxel + 1 other chemotherapy allowed + up to 2 NHA
50% PSA RR
ORR
Number of participants with AEs
Number of participants discontinuing study medication due to AEs
NCT02861573
1 BRCA1, BRCA2, ATM, PTEN, CHEK2, RAD51C, RAD51D, PALB2, MLH1, MSH2, MSH6, and PMS2; 2 Immunogenic signature: mismatch repair deficiency by IHC, defective DNA repair detected by a targeted sequencing panel, and high inflammatory infiltrate defined on multiplexed IHC criteria; 3 Composite response rate: radiological response (RECIST 1.1), PSA response ≥50% confirmed by a second PSA test at least 4 weeks later (PCWG3 2016), and conversion of CTC count from ≥5 cells/7.5 mL at baseline to <5 cells/7.5 mL confirmed by a second CTC test at least 4 weeks later (PCWG3 2016). 4 Immunogenic phenotype with of one of the next criteria: 1, mismatch repair deficiency and/or a high mutational burden of >7 mutations per Mb (cluster A); 2, BRCA2 inactivation or BRCAness signature (cluster B); 3, a tandem duplication signature and/or CDK12 biallelic inactivation (cluster C). 5 Disease control rate (DCR) of >6 months; this includes a change from baseline in tumor volume as measured by SD, PR, or CR by best ORR in evaluable participants, all lasting longer than 6 months; Abbreviations: mCRPC: metastatic castration resistant prostate cancer, mHNPC: metastatic hormone naïve prostate cancer; ns not specified AC: adenocarcinoma; t-NE: transdifferentiated neuroendocrine carcinoma of the prostate.
Table 3. Active trials with bispecific T cell engagers (BiTEs) and CAR-T cells in advanced PCa.
Table 3. Active trials with bispecific T cell engagers (BiTEs) and CAR-T cells in advanced PCa.
Short trial TitleTrial
Phase
Estimated
Enrolment
(pts)
Experimental TherapyDisease StageRequired PretreatmentPrimary EndpointNCT Number
Safety, Tolerability, Pharmacokinetics, and Efficacy of Acapatamab in Subjects With mCRPC1288Acapatamab,
acapatamab + Pembrolizumab,
acapatamab + Etanercept Prophylaxis,
acapatamab + Cytochrome P450 Cocktail
mCRPCADT, taxaneSafety and tolerabilityNCT03792841
A Study of Tarlatamab (AMG 757) in Participants with Neuroendocrine Prostate Cancer1b60 Tarlatamab (AMG 757)Neuroendocrine prostate cancer1 line of prior systemic treatmentSafety and tolerabilityNCT04702737
Study of AMG 509 in Subjects with Metastatic Castration-Resistant Prostate Cancer1110AMG 509mCRPCPrior NHA, taxaneSafety and tolerabilityNCT04221542
Safety and Efficacy of Therapies for Metastatic Castration-Resistant Prostate Cancer (mCRPC)1/2159Acapatamab + Enzalutamide,
Acapatamab + Abiraterone,
Acapatamab + AMG 404
mCRPC Safety and tolerabilityNCT04631601
Study with Bispecific Antibody Engaging T cells, in Patients with Progressive Cancer Diseases With Positive PSCA Marker124GEM3PSCAPSCA expressing cancer including prostate carcinomaProgressive Disease After Standard Systemic TherapyMTD
Incidence and intensity of AEs
DLT
NCT03927573
CART-PSMA-TGFβRDN Cells for Castrate-Resistant Prostate Cancer118CART-PSMA-TGFβRDNmCRPCAt least 1 NHASafety and tolerabilityNCT03089203
P-PSMA-101 CAR-T Cells in the Treatment of Subjects With mCRPC and Advanced Salivary Gland Cancers160P-PSMA-101
Rimiducid (safety switch activator) may be administered as indicated
mCRPC Safety, DLT, efficacy RECIST 1.1 and PCWG3NCT04249947
PSCA-CAR T Cells in Treating Patients with PSCA + mCRPC133Autologous Anti-PSCA-CAR-4-1BB/TCRzeta-CD19t-expressing T-lymphocytesmCRPCAt least 1 NHASafety and tolerability
Define recommended phase 2 dose
NCT03873805
Safety and Activity Study of PSCA-Targeted CAR-T Cells (BPX-601) in Subjects with Selected Advanced Solid Tumors1/2151BPX-601: Autologous T cells genetically modified with retrovirus vector containing PSCA-specific CAR and an inducible MyD88/Cluster designation (CD)40 (iMC) co-stimulatory domain
Rimiducid: Dimerizer infusion to activate the iMC of the BPX-601 cells for improved proliferation and persistence
mCRPC among others MTD and/or recommended extension dose of BPX-601 measured by DLTNCT02744287
A Study of JNJ-75229414 for Metastatic Castration-Resistant Prostate Cancer Participant160KLK2 CAR-T Cells (JNJ-75229414)mCRPCAt least 1 NHA or one prior chemotherapyNumber and severity of AE, DLTNCT05022849
Dose-Escalating Trial with UniCAR02-T Cells and PSMA Target Module (TMpPSMA) in Patients with Progressive Disease After Standard Systemic Therapy in Cancers With Positive PSMA Marker135UniCAR02-T Cells and PSMA Target Module (TMpPSMA)mCRPCSystemic standard therapiesSafety and tolerability, MTD, DLTNCT04633148
Abbreviations: MTD: maximum tolerated dose, AE: adverse event; DLT: dose-limiting toxicity.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

von Amsberg, G.; Alsdorf, W.; Karagiannis, P.; Coym, A.; Kaune, M.; Werner, S.; Graefen, M.; Bokemeyer, C.; Merkens, L.; Dyshlovoy, S.A. Immunotherapy in Advanced Prostate Cancer—Light at the End of the Tunnel? Int. J. Mol. Sci. 2022, 23, 2569. https://doi.org/10.3390/ijms23052569

AMA Style

von Amsberg G, Alsdorf W, Karagiannis P, Coym A, Kaune M, Werner S, Graefen M, Bokemeyer C, Merkens L, Dyshlovoy SA. Immunotherapy in Advanced Prostate Cancer—Light at the End of the Tunnel? International Journal of Molecular Sciences. 2022; 23(5):2569. https://doi.org/10.3390/ijms23052569

Chicago/Turabian Style

von Amsberg, Gunhild, Winfried Alsdorf, Panagiotis Karagiannis, Anja Coym, Moritz Kaune, Stefan Werner, Markus Graefen, Carsten Bokemeyer, Lina Merkens, and Sergey A. Dyshlovoy. 2022. "Immunotherapy in Advanced Prostate Cancer—Light at the End of the Tunnel?" International Journal of Molecular Sciences 23, no. 5: 2569. https://doi.org/10.3390/ijms23052569

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

Article Metrics

Back to TopTop