Preclinical Characterization and Phase I Trial Results of a Bispecific Antibody Targeting PD-L1 and 4-1BB (GEN1046) in Patients with Advanced Refractory Solid Tumors

Alexander Muik, Elena Garralda, Isil Altintas, Friederike Gieseke, Ravit Geva, Eytan Ben-Ami, Corinne Maurice-Dror, Emiliano Calvo, Patricia M LoRusso, Guzman Alonso, Maria E Rodriguez-Ruiz, Kristina B Schoedel, Jordan M Blum, Bianca Sänger, Theodora W Salcedo, Saskia M Burm, Eliana Stanganello, Dennis Verzijl, Fulvia Vascotto, Angelica Sette, Juliane Quinkhardt, Theo S Plantinga, Aras Toker, Edward N van den Brink, Mark Fereshteh, Mustafa Diken, David Satijn, Sebastian Kreiter, Esther C W Breij, Gaurav Bajaj, Eleni Lagkadinou, Kate Sasser, Özlem Türeci, Ulf Forssmann, Tahamtan Ahmadi, Uğur Şahin, Maria Jure-Kunkel, Ignacio Melero, Alexander Muik, Elena Garralda, Isil Altintas, Friederike Gieseke, Ravit Geva, Eytan Ben-Ami, Corinne Maurice-Dror, Emiliano Calvo, Patricia M LoRusso, Guzman Alonso, Maria E Rodriguez-Ruiz, Kristina B Schoedel, Jordan M Blum, Bianca Sänger, Theodora W Salcedo, Saskia M Burm, Eliana Stanganello, Dennis Verzijl, Fulvia Vascotto, Angelica Sette, Juliane Quinkhardt, Theo S Plantinga, Aras Toker, Edward N van den Brink, Mark Fereshteh, Mustafa Diken, David Satijn, Sebastian Kreiter, Esther C W Breij, Gaurav Bajaj, Eleni Lagkadinou, Kate Sasser, Özlem Türeci, Ulf Forssmann, Tahamtan Ahmadi, Uğur Şahin, Maria Jure-Kunkel, Ignacio Melero

Abstract

Checkpoint inhibitors (CPI) have revolutionized the treatment paradigm for advanced solid tumors; however, there remains an opportunity to improve response rates and outcomes. In preclinical models, 4-1BB costimulation synergizes with CPIs targeting the programmed cell death protein 1 (PD-1)/programmed cell death ligand 1 (PD-L1) axis by activating cytotoxic T-cell-mediated antitumor immunity. DuoBody-PD-L1×4-1BB (GEN1046) is an investigational, first-in-class bispecific immunotherapy agent designed to act on both pathways by combining simultaneous and complementary PD-L1 blockade and conditional 4-1BB stimulation in one molecule. GEN1046 induced T-cell proliferation, cytokine production, and antigen-specific T-cell-mediated cytotoxicity superior to clinically approved PD-(L)1 antibodies in human T-cell cultures and exerted potent antitumor activity in transplantable mouse tumor models. In dose escalation of the ongoing first-in-human study in heavily pretreated patients with advanced refractory solid tumors (NCT03917381), GEN1046 demonstrated pharmacodynamic immune effects in peripheral blood consistent with its mechanism of action, manageable safety, and early clinical activity [disease control rate: 65.6% (40/61)], including patients resistant to prior PD-(L)1 immunotherapy.

Significance: DuoBody-PD-L1×4-1BB (GEN1046) is a first-in-class bispecific immunotherapy with a manageable safety profile and encouraging preclinical and early clinical activity. With its ability to confer clinical benefit in tumors typically less sensitive to CPIs, GEN1046 may fill a clinical gap in CPI-relapsed or refractory disease or as a combination therapy with CPIs. See related commentary by Li et al., p. 1184. This article is highlighted in the In This Issue feature, p. 1171.

©2022 The Authors; Published by the American Association for Cancer Research.

Figures

Figure 1.
Figure 1.
GEN1046 induces dose-dependent, conditional T-cell proliferation and cytokine production and enhances antigen-specific T-cell–mediated cytotoxicity in vitro. A and B, Activated T cells were cocultured with autologous iDCs in the presence of GEN1046 or control antibodies (0.125 µg/mL), and T-cell/iDC clusters were visualized over time by live-cell imaging. Quantification of the number of T cells in contact with a given DC on average (A), or over time (B; left), as well as the duration of these DC/T-cell clusters (B; right) is shown. ****, P < 0.0001; ns, not significant; Mann–Whitney U test. C, Induction of 4-1BB signaling by GEN1046 or control antibodies was assessed using a 4-1BB reporter assay. D, Blockade of the PD-1/PD-L1 interaction by GEN1046 or control antibodies was assessed using a PD-1/PD-L1 blockade bioassay. E, CFSE-labeled human PBMCs were stimulated with anti-CD3 (0.1 µg/mL) and incubated with GEN1046 or control antibodies (0.2 µg/mL) for 4 days. CFSE dilution in CD8+ T cells was analyzed by flow cytometry, and the expansion index was calculated. Data shown are the fold change in expansion index of treatment groups, relative to untreated cells of individual donors, as well as mean ± standard deviation (n = 11). ****, P < 0.0001; **, P < 0.01; *, P < 0.05, Friedman test with Dunn multiple comparisons test. F, CD8+ T cells were electroporated with RNA encoding a CLDN6-specific TCR and PD-1, labeled with CFSE, and cocultured with autologous DCs electroporated with CLDN6-encoding RNA in the presence of GEN1046 or control antibodies (0.2 µg/mL) for 4 days. Proliferation was measured by CFSE dilution as described in E. ****, P < 0.0001, Friedman test with Dunn multiple comparisons test. G, IFNγ concentrations in supernatant taken after 48 hours from cultures as described in F. Data shown are mean concentration ± standard deviation of triplicate wells from one representative donor (n = 3 donors). H–K, CD8+ T cells were electroporated with RNA encoding a CLDN6-specific TCR and were preactivated for 24 hours in coculture with CLDN6-expressing MDA-MB-231 cells (MDA-MB-231 hCLDN6) to induce 4-1BB expression. Subsequently, the preactivated CD8+ T cells were transferred to cocultures with previously seeded MDA-MB-231 hCLDN6 tumor cells (hCLDN6+ PD-L1+) in the presence of GEN1046 or control antibodies (0.2 µg/mL). HI, After 48 hours, expression of cytotoxic mediator GZMB and degranulation marker CD107a by the CD8+ T cells was assessed; pooled data from four experiments (n = 3–12) are depicted. MFIs are expressed as relative values compared with the isotype control condition. J, Cytotoxicity of the CD8+ T cells toward MDA-MB-231 CLDN6 cells was monitored by electrical impedance measurement over 5 days using the xCELLigence real-time cell analysis. Representative data (mean ± standard deviation of triplicate wells) from cocultures derived from one individual donor are shown (n = 12). Data were normalized to the time point of coculture start and expressed relative to tumor-cell cultures without T cells (without T cells set to 100%). K, AUC (total area) analysis of cytotoxicity data (n = 11–12). ****, P < 0.0001; ***, P < 0.001; **, P < 0.01; *, P < 0.05, mixed-effect analysis with Dunnett multiple comparisons test. MFI, median fluorescence intensity; ns, not significant.
Figure 2.
Figure 2.
GEN1046 therapeutic efficacy and antitumor immune responses in MC38-hPD-L1 tumor–bearing hPD-L1/h4-1BB dKI mice. A, MC38-hPD-L1 cells (1 × 106 tumor cells) were injected s.c. in the right flank of hPD-L1/h4-1BB dKI mice. After tumor establishment (average tumor volume, ∼80 mm3), mice were randomized and treated with GEN1046 or isotype control antibody (each 5 mg/kg intravenous) at the indicated time points (n = 9 per group). B, Tumor growth of individual mice in each group, with CR defined as number of animals with CR. C, Progression-free survival, defined as the percentage of mice with tumor volume smaller than 500 mm3, is shown as a Kaplan–Meier curve. Mantel–Cox analysis was used to compare survival between treatment groups. ****, P < 0.0001. D, Mice with CR after treatment with GEN1046 (shown in B) were rechallenged by s.c. injection of 1 × 106 MC38-hPD-L1 tumor cells in the left flank on day 164. As a control group, a second cohort of naïve dKI animals was inoculated with 1 × 106 MC38-hPD-L1 tumor cells. Tumor growth of individual mice in each group is shown. E–I, MC38-hPD-L1 or MC38-WT cells (1 × 106 tumor cells) were injected s.c. in the right flank of hPD-L1/h4-1BB dKI mice. After tumor establishment (mean tumor volume, ∼75 mm3), mice were randomized and treated with GEN1046, a durvalumab analogue, or isotype control antibody (all 5 mg/kg intravenous) at the indicated time points. The mice were sacrificed 2 days after the last treatment (n = 5 per treatment group), and the tumors and spleens were excised. F, Sections of resected tumors (4 µm) were stained using anti-CD3, anti-CD8, or anti-FoxP3 antibodies by IHC. The number of positive cells was quantified per mm2. Mann–Whitney U statistical analysis was performed to compare the number of cellular subsets between treatment groups in MC38-hPD-L1 or MC38-WT tumor–bearing mice. *, P < 0.05. G and H, Flow cytometry analysis of dissociated splenocytes. Data from individual mice are shown as well as group mean ± SEM. *, P < 0.05; **, P < 0.01; ***, P < 0.001, Wilcoxon rank sum test. I, Peripheral blood samples were taken 1 day before treatment (d−1) and 2 days after each treatment. Cytokine analysis was performed by electrochemiluminescence immunoassay. s.c., subcutaneous; Treg, regulatory T cell.
Figure 3.
Figure 3.
GEN1046 promotes TIL expansion from patient-derived tumor tissue. Tumor tissues resected from patients with NSCLC were cut into pieces of 1 to 2 mm3 and cultured in the presence of IL2 (10–50 U/mL) and GEN1046 (or a GEN1046 surrogate comprising the PD-L1–specific Fab arm of GEN1046 and a nonhumanized variant of the 4-1BB–specific Fab arm or atezolizumab (0.2 µg/mL), or with IL2 only for 14–17 days. A, Cell numbers after expansion were determined by flow cytometry, and total TILs, CD8+ T cells, CD4+ T cells, and NK cells are shown for three patients. Tumor PD-L1 expression and 4-1BB expression by CD8+ T cells in the specimen at baseline are indicated. B, TCR repertoire analysis was performed by TRB RNA sequencing of the expanded TILs and the tumor fragments. Cumulative frequency of shared clonotypes, the 20 most abundant clonotypes in the GEN1046 surrogate–treated cultures, is shown. C, TILs expanded as in A were restimulated with enzymatically digested autologous tumor (TC) in the presence or absence of an MHC I–blocking antibody. Expression of 4-1BB and intracellular expression of IFNγ and CD107a in CD8+ T cells were analyzed by flow cytometry.
Figure 4.
Figure 4.
Pharmacokinetics and pharmacodynamics of GEN1046 in patients with advanced solid tumors. A, Mean plasma concentration of GEN1046 during the first two dosing cycles with administration Q3W. B and C, Maximal fold change from baseline in pharmacodynamic markers measured in peripheral blood during cycle 1 in patients receiving low (≤200 mg) and high (≥400 mg) doses of GEN1046. P values from the Wilcoxon–Mann–Whitney test. LLOQ, lower limit of quantification.
Figure 5.
Figure 5.
GEN1046 antitumor efficacy in patients with advanced solid tumors. A, Waterfall plot of best relative percent change from baseline in tumor size. BE, CT over time (B and C) and plots of immunologic changes in the periphery (D and E) following the first dose of GEN1046 in a patient with metastatic NSCLC treated with 200 mg GEN1046 (B and D) and a patient with metastatic ovarian cancer treated with 80 mg GEN1046 (C and E). Numbers (percentages) below CT scans are sum of diameter of target lesion (percent change from baseline). NA, not applicable; NE, not evaluable. aArchival formalin-fixed paraffin-embedded block of tumor tissue obtained >5 years prior was evaluated; loss of PD-L1 immunoreactivity may have occurred.

References

    1. Nixon NA, Blais N, Ernst S, Kollmannsberger C, Bebb G, Butler M, et al. . Current landscape of immunotherapy in the treatment of solid tumours, with future opportunities and challenges. Curr Oncol 2018;25:e373–e84.
    1. Sharma P, Siddiqui BA, Anandhan S, Yadav SS, Subudhi SK, Gao J, et al. . The next decade of immune checkpoint therapy. Cancer Discov 2021;11:838–57.
    1. Jenkins RW, Barbie DA, Flaherty KT. Mechanisms of resistance to immune checkpoint inhibitors. Br J Cancer 2018;118:9–16.
    1. Schadendorf D, Hodi FS, Robert C, Weber JS, Margolin K, Hamid O, et al. . Pooled analysis of long-term survival data from phase II and phase III trials of ipilimumab in unresectable or metastatic melanoma. J Clin Oncol 2015;33:1889–94.
    1. Sun JY, Zhang D, Wu S, Xu M, Zhou X, Lu XJ, et al. . Resistance to PD-1/PD-L1 blockade cancer immunotherapy: mechanisms, predictive factors, and future perspectives. Biomark Res 2020;8:35.
    1. Maleki Vareki S, Garrigós C, Duran I. Biomarkers of response to PD-1/PD-L1 inhibition. Crit Rev Oncol Hematol 2017;116:116–24.
    1. Gong J, Chehrazi-Raffle A, Reddi S, Salgia R. Development of PD-1 and PD-L1 inhibitors as a form of cancer immunotherapy: a comprehensive review of registration trials and future considerations. J Immunother Cancer 2018;6:8.
    1. Betof Warner A, Palmer JS, Shoushtari AN, Goldman DA, Panageas KS, Hayes SA, et al. . Long-term outcomes and responses to retreatment in patients with melanoma treated with PD-1 blockade. J Clin Oncol 2020;38:1655–63.
    1. Chester C, Sanmamed MF, Wang J, Melero I. Immunotherapy targeting 4-1BB: mechanistic rationale, clinical results, and future strategies. Blood 2018;131:49–57.
    1. Shuford WW, Klussman K, Tritchler DD, Loo DT, Chalupny J, Siadak AW, et al. . 4-1BB costimulatory signals preferentially induce CD8+ T cell proliferation and lead to the amplification in vivo of cytotoxic T cell responses. J Exp Med 1997;186:47–55.
    1. Pollok KE, Kim YJ, Zhou Z, Hurtado J, Kim KK, Pickard RT, et al. . Inducible T cell antigen 4-1BB: analysis of expression and function. J Immunol 1993;150:771–81.
    1. Horton BL, Williams JB, Cabanov A, Spranger S, Gajewski TF. Intratumoral CD8(+) T-cell apoptosis is a major component of T-cell dysfunction and impedes antitumor immunity. Cancer Immunol Res 2018;6:14–24.
    1. Melero I, Shuford WW, Newby SA, Aruffo A, Ledbetter JA, Hellström KE, et al. . Monoclonal antibodies against the 4-1BB T-cell activation molecule eradicate established tumors. Nat Med 1997;3:682–5.
    1. Hosoi A, Takeda K, Nagaoka K, Iino T, Matsushita H, Ueha S, et al. . Increased diversity with reduced "diversity evenness" of tumor infiltrating T-cells for the successful cancer immunotherapy. Sci Rep 2018;8:1058.
    1. Reithofer M, Rosskopf S, Leitner J, Battin C, Bohle B, Steinberger P, et al. . 4-1BB costimulation promotes bystander activation of human CD8 T cells. Eur J Immunol 2021;51:721–33.
    1. Gros A, Robbins PF, Yao X, Li YF, Turcotte S, Tran E, et al. . PD-1 identifies the patient-specific CD8+ tumor-reactive repertoire infiltrating human tumors. J Clin Invest 2014;124:2246–59.
    1. Ye Q, Song DG, Poussin M, Yamamoto T, Best A, Li C, et al. . CD137 accurately identifies and enriches for naturally occurring tumor-reactive T cells in tumor. Clin Cancer Res 2014;20:44–55.
    1. Etxeberria I, Glez-Vaz J, Teijeira Á, Melero I. New emerging targets in cancer immunotherapy: CD137/4-1BB costimulatory axis. ESMO Open 2020;4:e000733.
    1. Bartkowiak T, Curran MA. 4-1BB agonists: multi-potent potentiators of tumor immunity. Front Oncol 2015;5:117.
    1. Ascierto PA, Simeone E, Sznol M, Fu YX, Melero I. Clinical experiences with anti-CD137 and anti–PD-1 therapeutic antibodies. Semin Oncol 2010;37:508–16.
    1. Timmerman J, Herbaux C, Ribrag V, Zelenetz AD, Houot R, Neelapu SS, et al. . Urelumab alone or in combination with rituximab in patients with relapsed or refractory B-cell lymphoma. Am J Hematol 2020;95:510–20.
    1. Segal NH, Logan TF, Hodi FS, McDermott D, Melero I, Hamid O, et al. . Results from an integrated safety analysis of urelumab, an agonist anti-CD137 monoclonal antibody. Clin Cancer Res 2017;23:1929–36.
    1. Segal NH, He AR, Doi T, Levy R, Bhatia S, Pishvaian MJ, et al. . Phase I study of single-agent utomilumab (PF-05082566), a 4-1BB/CD137 agonist, in patients with advanced cancer. Clin Cancer Res 2018;24:1816–23.
    1. Labrijn AF, Meesters JI, de Goeij BE, van den Bremer ET, Neijssen J, van Kampen MD, et al. . Efficient generation of stable bispecific IgG1 by controlled Fab-arm exchange. Proc Natl Acad Sci U S A 2013;110:5145–50.
    1. Labrijn AF, Meesters JI, Priem P, de Jong RN, van den Bremer ET, van Kampen MD, et al. . Controlled Fab-arm exchange for the generation of stable bispecific IgG1. Nat Protoc 2014;9:2450–63.
    1. Dustin ML. The immunological synapse. Cancer Immunol Res 2014;2:1023–33.
    1. Betts A, van der Graaf PH. Mechanistic quantitative pharmacology strategies for the early clinical development of bispecific antibodies in oncology. Clin Pharmacol Ther 2020;108:528–41.
    1. Bajaj G, Nazari F, Presler M, Thalhauser C, Forssmann U, Jure-Kunkel M, et al. . Dose selection for DuoBody-PD-L1×4-1BB (GEN1046) using a semimechanistic pharmacokinetics/pharmacodynamics model that leverages preclinical and clinical data [abstract 786]. J Immunother Cancer 2021;9:A821.
    1. Eisenhauer EA, Therasse P, Bogaerts J, Schwartz LH, Sargent D, Ford R, et al. . New response evaluation criteria in solid tumours: revised RECIST guideline (version 1.1). Eur J Cancer 2009;45:228–47.
    1. Simoni Y, Becht E, Fehlings M, Loh CY, Koo SL, Teng KWW, et al. . Bystander CD8(+) T cells are abundant and phenotypically distinct in human tumour infiltrates. Nature 2018;557:575–9.
    1. Clouthier DL, Lien SC, Yang SYC, Nguyen LT, Manem VSK, Gray D, et al. . An interim report on the investigator-initiated phase 2 study of pembrolizumab immunological response evaluation (INSPIRE). J Immunother Cancer 2019;7:72.
    1. Teng MW, Ngiow SF, Ribas A, Smyth MJ. Classifying cancers based on T-cell infiltration and PD-L1. Cancer Res 2015;75:2139–45.
    1. Wculek SK, Cueto FJ, Mujal AM, Melero I, Krummel MF, Sancho D. Dendritic cells in cancer immunology and immunotherapy. Nat Rev Immunol 2020;20:7–24.
    1. Qu QX, Zhu XY, Du WW, Wang HB, Shen Y, Zhu YB, et al. . 4-1BB agonism combined with PD-L1 blockade increases the number of tissue-resident CD8+ T cells and facilitates tumor abrogation. Front Immunol 2020;11:577.
    1. Azpilikueta A, Agorreta J, Labiano S, Pérez-Gracia JL, Sánchez-Paulete AR, Aznar MA, et al. . Successful immunotherapy against a transplantable mouse squamous lung carcinoma with anti–PD-1 and anti-CD137 monoclonal antibodies. J Thorac Oncol 2016;11:524–36.
    1. Seifert AM, Eymer A, Heiduk M, Wehner R, Tunger A, von Renesse J, et al. . PD-1 expression by lymph node and intratumoral regulatory T cells Is associated with lymph node metastasis in pancreatic cancer. Cancers 2020;12:2756.
    1. Lakins MA, Koers A, Giambalvo R, Munoz-Olaya J, Hughes R, Goodman E, et al. . FS222, a CD137/PD-L1 tetravalent bispecific antibody, exhibits low toxicity and antitumor activity in colorectal cancer models. Clin Cancer Res 2020;26:4154–67.
    1. Zhai T, Wang C, Xu Y, Huang W, Yuan Z, Wang T, et al. . Generation of a safe and efficacious llama single-domain antibody fragment (vHH) targeting the membrane-proximal region of 4-1BB for engineering therapeutic bispecific antibodies for cancer. J Immunother Cancer 2021;9:e002131.
    1. Jeong S, Park E, Kim HD, Sung E, Kim H, Jeon J, et al. . Novel anti-4-1BB×PD-L1 bispecific antibody augments anti-tumor immunity through tumor-directed T-cell activation and checkpoint blockade. J Immunother Cancer 2021;9:e002428.
    1. Geuijen C, Tacken P, Wang LC, Klooster R, van Loo PF, Zhou J, et al. . A human CD137×PD-L1 bispecific antibody promotes anti-tumor immunity via context-dependent T cell costimulation and checkpoint blockade. Nat Commun 2021;12:4445.
    1. Qiao Y, Qiu Y, Ding J, Luo N, Wang H, Ling X, et al. . Cancer immune therapy with PD-1-dependent CD137 co-stimulation provides localized tumour killing without systemic toxicity. Nat Commun 2021;12:6360.
    1. Kelley RK, Sangro B, Harris W, Ikeda M, Okusaka T, Kang YK, et al. . Safety, efficacy, and pharmacodynamics of tremelimumab plus durvalumab for patients with unresectable hepatocellular carcinoma: randomized expansion of a phase I/II study. J Clin Oncol 2021;39:2991–3001.
    1. Yamazaki N, Kiyohara Y, Uhara H, Iizuka H, Uehara J, Otsuka F, et al. . Cytokine biomarkers to predict antitumor responses to nivolumab suggested in a phase 2 study for advanced melanoma. Cancer Sci 2017;108:1022–31.
    1. Kamphorst AO, Pillai RN, Yang S, Nasti TH, Akondy RS, Wieland A, et al. . Proliferation of PD-1+ CD8 T cells in peripheral blood after PD-1-targeted therapy in lung cancer patients. Proc Natl Acad Sci U S A 2017;114:4993–8.
    1. Huang AC, Postow MA, Orlowski RJ, Mick R, Bengsch B, Manne S, et al. . T-cell invigoration to tumour burden ratio associated with anti–PD-1 response. Nature 2017;545:60–5.
    1. Heinisch IV, Daigle I, Knöpfli B, Simon HU. CD137 activation abrogates granulocyte-macrophage colony-stimulating factor-mediated anti-apoptosis in neutrophils. Eur J Immunol 2000;30:3441–6.
    1. Niu L, Strahotin S, Hewes B, Zhang B, Zhang Y, Archer D, et al. . Cytokine-mediated disruption of lymphocyte trafficking, hemopoiesis, and induction of lymphopenia, anemia, and thrombocytopenia in anti-CD137-treated mice. J Immunol 2007;178:4194–213.
    1. Muik A, Altinas I, Gieseke F, Schoedel KB, Burm SM, Toker A, et al. . An Fc-inert PD-L1×4-1BB bispecific antibody mediates potent anti-tumor immunity in mice by combining checkpoint inhibition and conditional 4-1BB co-stimulation. Oncoimmunology 2022;11:2030135.
    1. Vink T, Oudshoorn-Dickmann M, Roza M, Reitsma JJ, de Jong RN. A simple, robust and highly efficient transient expression system for producing antibodies. Methods 2014;65:5–10.
    1. Engelberts PJ, Hiemstra IH, de Jong B, Schuurhuis DH, Meesters J, Beltran Hernandez I, et al. . DuoBody-CD3xCD20 induces potent T-cell-mediated killing of malignant B cells in preclinical models and provides opportunities for subcutaneous dosing. EBioMedicine 2020;52:102625.
    1. Barbas CF III, Collet TA, Amberg W, Roben P, Binley JM, Hoekstra D, et al. . Molecular profile of an antibody response to HIV-1 as probed by combinatorial libraries. J Mol Biol 1993;230:812–23.
    1. Faries D. Practical modifications of the continual reassessment method for phase I cancer clinical trials. J Biopharm Stat 1994;4:147–64.
    1. Babb J, Rogatko A, Zacks S. Cancer phase I clinical trials: efficient dose escalation with overdose control. Stat Med 1998;17:1103–20.
    1. Seymour L, Bogaerts J, Perrone A, Ford R, Schwartz LH, Mandrekar S, et al. . iRECIST: guidelines for response criteria for use in trials testing immunotherapeutics. Lancet Oncol 2017;18:e143–e52.

Source: PubMed

3
Tilaa