Next Article in Journal
Genes and Longevity of Lifespan
Next Article in Special Issue
Resistance Mechanisms in Pediatric B-Cell Acute Lymphoblastic Leukemia
Previous Article in Journal
Repolarization of Unbalanced Macrophages: Unmet Medical Need in Chronic Inflammation and Cancer
Previous Article in Special Issue
Melittin Increases Cisplatin Sensitivity and Kills KM-H2 and L-428 Hodgkin Lymphoma Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Aspects of Resistance to Immunotherapies—Advances in Understanding and Management of Diffuse Large B-Cell Lymphoma

by
Aleksandra Kusowska
1,2,
Matylda Kubacz
1,
Marta Krawczyk
1,3,4,
Aleksander Slusarczyk
1,5,
Magdalena Winiarska
1,3 and
Malgorzata Bobrowicz
1,*
1
Department of Immunology, Medical University of Warsaw, 02-097 Warsaw, Poland
2
Doctoral School, Medical University of Warsaw, 02-091 Warsaw, Poland
3
Laboratory of Immunology, Mossakowski Medical Research Institute, Polish Academy of Sciences, 02-106 Warsaw, Poland
4
Doctoral School of Translational Medicine, Centre of Postgraduate Medical Education, 01-813 Warsaw, Poland
5
Department of General, Oncological and Functional Urology, Medical University of Warsaw, 02-005 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(3), 1501; https://doi.org/10.3390/ijms23031501
Submission received: 22 December 2021 / Revised: 22 January 2022 / Accepted: 26 January 2022 / Published: 28 January 2022
(This article belongs to the Special Issue Deciphering the Therapeutic Resistance of Hematologic Malignancies)

Abstract

:
Despite the unquestionable success achieved by rituximab-based regimens in the management of diffuse large B-cell lymphoma (DLBCL), the high incidence of relapsed/refractory disease still remains a challenge. The widespread clinical use of chemo-immunotherapy demonstrated that it invariably leads to the induction of resistance; however, the molecular mechanisms underlying this phenomenon remain unclear. Rituximab-mediated therapeutic effect primarily relies on complement-dependent cytotoxicity and antibody-dependent cell cytotoxicity, and their outcome is often compromised following the development of resistance. Factors involved include inherent genetic characteristics and rituximab-induced changes in effectors cells, the role of ligand/receptor interactions between target and effector cells, and the tumor microenvironment. This review focuses on summarizing the emerging advances in the understanding of the molecular basis responsible for the resistance induced by various forms of immunotherapy used in DLBCL. We outline available models of resistance and delineate solutions that may improve the efficacy of standard therapeutic protocols, which might be essential for the rational design of novel therapeutic regimens.

1. Introduction

1.1. Diffuse Large B-Cell Lymphoma

Non-Hodgkin lymphoma (NHL) is the most frequent hematological neoplasm in the world [1] with more than 544,000 new NHL cases diagnosed in 2020 (2.8% of all cancer diagnoses) [2]. Of all the NHL subtypes, the most common is diffuse large B-cell lymphoma (DLBCL), accounting for approximately 40% of lymphoma cases. DLBCL is also one of the most aggressive subtypes; 5-year survival in elderly patients does not exceed 40% [3].
The most common first-line treatment for DLBCL is chemoimmunotherapy containing rituximab, the so-called R-CHOP regimen (rituximab, cyclophosphamide, doxorubicin, vincristine), which fails in 30–40% of patients. Although relapsed/refractory (r/r) patients receive second-line treatment or may undergo autologous stem cell transplantation, their prognosis remains poor [4]. As demonstrated by the SCHOLAR-1 trial, the objective response rate (ORR) of r/r patients treated with next-line therapy was 26%, and the median overall survival was 6.3 months [5]. The greatest challenge in DLBCL therapy is the genetic diversity of this neoplasm, i.e., the presence of mutations, translocations, and chromosomal aberrations. In the face of this heterogeneity, investigation of novel common antigens, which can be targeted by immunotherapy and identified by a universal approach to eliminate malignant cells regardless of their molecular pathogenesis, is constantly pursued [6].
Recently, r/r patients have gained the access to adoptive therapy using genetically engineered T cells with chimeric antigen receptor (CAR-T cells) targeting DLBCL-expressed surface antigens. Particularly promising results of this treatment strategy have been observed in the case of CD19-CARs. Three of them, i.e., axicabtagene ciloleucel (Yescarta, Kite), lisocabtagene maraleucel (Breyanzi, Bristol Myers Squibb), and tisagenlecleucel (Kymriah, Novartis) are licensed to treat DLBCL patients [7].
However, years of employment of rituximab as a core of first-line treatment and recent observations on CAR-T cell therapy clearly demonstrate that the use of immunotherapy invariably leads to the induction of resistance [8]. A deep understanding of the molecular mechanisms underlying this phenomenon is essential for the rational design of novel therapeutic regimens. Therefore, this review aims at summarizing the mechanisms of resistance induced by various forms of immunotherapy used in DLBCL (Figure 1).

1.2. Rituximab (RTX)

Targeting CD20 antigen using rituximab, a monoclonal antibody (mAb) approved by the FDA in 1997, remains the most widely used form of immunotherapy in oncology. CD20 is a membrane protein found on the surface of normal and neoplastic B lymphocytes, but not on hematopoietic progenitors, which makes it a suitable target for mAbs [9]. In clinical trials, rituximab combined with chemotherapy significantly improved progression-free survival (PFS) and overall survival (OS) rates in B-cell-derived malignancies such as chronic lymphocytic leukemia (CLL), follicular lymphoma (FL), and DLBCL [10]. Experience with rituximab provided proof of concept that mAbs could be used in cancer treatment and the combination of rituximab with chemotherapy, i.e., the R-CHOP regimen has remained the first-line of treatment for DLBCL patients for years.

1.3. Mechanisms of Rituximab Cytotoxicity

The cytotoxic effect of rituximab is exerted through several mechanisms, relying on the activation of anti-tumor immune response (reviewed in [11,12]). These include complement-dependent cytotoxicity (CDC), antibody-dependent cellular cytotoxicity (ADCC), antibody-dependent cellular phagocytosis (ADCP), and direct apoptosis induction [13]. Additionally, rituximab sensitizes non-Hodgkin lymphoma cells to chemotherapy [14,15]. Rituximab hampers the secretion of interleukin (IL)-10, which is produced by NHL cells as a protective factor. IL-10 induces STAT3 phosphorylation and in consequence the expression of an anti-apoptotic protein Bcl-2, which leads to the resistance to chemotherapeutics [14,15]. Moreover, by upregulating Raf-1 kinase inhibitor protein (RKIP), rituximab decreases the signaling of the ERK1/2 and NF-kappa B pathways, leading to downregulation of Bcl-xL [15]. Furthermore, in NHL cell lines, rituximab inhibits the overexpressed transcription repressor Yin Yang 1 (YY1), which suppresses Fas and DR5 expression, leading to sensitization to Fas ligand and tumor necrosis factor (TNF)-related apoptosis-inducing ligand (TRAIL)-induced apoptosis [16].
The results of preclinical studies in vitro and in vivo in animal models as well as observations from the clinical settings suggest that all the above-mentioned mechanisms contribute to the overall efficacy of rituximab. However, it seems that the influence of single effector mechanisms of rituximab may vary depending on the tumor type.

2. Resistance to Antibody-Dependent Cellular Cytotoxicity

It has been proposed by several studies that ADCC is the leading mechanism of action of rituximab in lymphomas [17,18,19,20]. ADCC relies on the recognition of Fc fragment of the antibody that opsonizes the target cell by the Fc receptor (FcR) present on the surface of the effector cell, which in turn undergoes degranulation. It has been demonstrated that natural killer (NK) cells are the major effectors of ADCC [21]. This effect is predominantly exerted by a subpopulation of CD56dim NK cells with higher levels of FcγRIII receptor (CD16) [22]. Due to their ability to mount a rapid immune response and no need for priming, they play a crucial role in initiating anti-tumor response [23]. By now, there is a strong body of evidence describing various factors shaping the primary resistance and warranting the worst prognosis. Thus far, neither in vitro nor in vivo models studying the development of resistance through rituximab-induced ADCC have been generated.

2.1. Mechanisms of Primary Resistance to ADCC

2.1.1. Effector Cell Characteristics Influencing ADCC Efficacy

Genetic Polymorphism in FcγRIII Receptor

The efficiency of NK cell-mediated ADCC is highly influenced by the ability of FcγRIII receptor to recognize Fc portion of immunoglobulins [24]. Polymorphisms in the gene encoding FcγRIIIa (CD16a) determine response rates to R-CHOP treatment. The polymorphism at nucleotide position 559 (G or T) leads to the synthesis of valine (V) or phenylalanine (F), respectively. This in turn results in three different genotypes: homozygous FcγRIIIa-158 F/F or FcγRIIIa-158 V/V, and heterozygous FcγRIIIa-158 V/F, each characterized by diverse binding affinity to Fc portion of mAbs, and thus the variable rate of NK-cell cytotoxicity [25]. Among them, V/V genotype was described as high-affinity, whereas F/F as a low-affinity variant [26]. Several studies demonstrated the advantage of V/V over F/F genotype in association with the response to rituximab and R-CHOP regimen [27,28], but also with PFS and OS in large cohorts of DLBCL patients following R-CHOP treatment [29,30,31], especially in germinal center-derived (GCB) subset of DLBCL patients [32]. On the contrary, some studies have questioned the importance of FcγRIIIa polymorphism in mediating effective ADCC response to R-CHOP [33] or its role in predicting prognosis in DLBCL [34], which suggests the existence of other factors influencing the efficiency of ADCC in tumor eradication.

Low CD16 Expression

One of the reasons for the impaired ADCC activity in DLBCL patients might be an inherently low expression of CD16 on effector cells. A study on 36 newly diagnosed DLBCL patients presented markedly decreased protein levels of CD16 and CD137, a co-stimulatory molecule expressed on activated NK cells, as compared to healthy controls. This resulted in impaired degranulation in the presence of RTX, which could not be rescued by stimulation with IL-2 [35]. A comprehensive analysis of phenotypic and functional characteristics of NK cells at DLBCL diagnosis followed by RTX treatment revealed that the initial high level of NK CD16-positive cells including CD56dim population decreased over the course of therapy, whereas the expression of activating receptors (CD16, NKG2D) was diminished [36].

KIR/HLA Interactions

ADCC function of NK cells is strongly influenced by the interactions between inhibitory killer cell Ig–like receptors (KIRs) and human leukocyte antigen (HLA) class I on target cells [37]. The binding of some KIR ligands to KIRs hampers NK-cell degranulation [38]. A study on NK cells isolated from the blood of healthy donors revealed that the interaction of KIR2DL1/KIR2DL2/KIR2DL3 with HLA-C and KIR3DL1 with HLA-Bw4 on EBV-B cell line significantly reduces rituximab-mediated degranulation of educated NK cells and ADCC efficacy [39]. Further, investigation of clinical response to rituximab in 74 NHL patients demonstrated that low KIR ligand number and high frequency of KIR-positive NK cells, which are notably efficient in driving RTX-mediated ADCC, correlated with improved complete response (CR) and prognosis. Thus, it has been postulated that low KIR ligand number and peripheral KIR-negative NK cells are good prognostic markers of response to RTX-based regimens [39].
However, to the best of our knowledge, it has not been demonstrated thus far whether lymphoma cells modulate the expression of KIR ligands on their surface or induce changes in the expression of KIRs on the surface of effector cells following rituximab treatment.

2.1.2. The Influence of the Microenvironment on ADCC Efficacy

Recent findings indicate that the presence of other immune cells may affect ADCC potency. Depletion of NK cells from peripheral blood mononuclear cells (PBMCs) entirely abrogated ADCC, whereas removal of monocytes increased ADCC efficacy [40]. Further, pre-incubation of rituximab-coated lymphoma cells with THP-1 monocyte cell line resulted in the suppression of NK cell-mediated ADCC [40] since monocytes were found to remove rituximab from the cell surface of B-cells [41]. The phenomenon of monocyte-mediated shaving of RTX/CD20 complexes was found to rely on serine protease activity since protease inhibitors can partially suppress RTX removal from the cell surface [42].
Monocytes and neutrophils can produce NOX2-derived reactive oxygen species (ROS) known to play immunomodulatory functions [43]. NK cells were found to be particularly sensitive to oxidative stress induced by hydrogen peroxide (H2O2), which markedly decreased NK-cell viability and significantly impaired the ability to mediate ADCC. The reduced anti-tumor potential of NK cells was correlated with intrinsically low expression of peroxiredoxin-1 (PRDX1), whereas its overexpression improved NK-cell function [44].
Anti-CD20 mAbs promote the release of ROS from neutrophils in CLL patient samples [45] and lymphoma cell lines [46]. Rituximab and ofatumumab were shown to trigger monocytes to generate ROS in a NOX-2 mediated pathway by attaching to Fc portion of anti-CD20 mAb [47]. In turn, NK-cell mediated ADCC against primary CLL cells was markedly reduced in the presence of monocytes and only partial rescue of ADCC function occurred in presence of anti-oxidants suggesting the involvement of more complex mechanisms associated with monocyte-mediated ADCC inhibition [47].

2.1.3. Characteristics of Malignant B-Cells Contributing to Impaired ADCC

It is widely accepted that tumor cells develop strategies to avoid recognition and eradication by the immune system. This is also true for DLBCL cells [48,49]. Data from The Leukemia and Lymphoma Molecular Profiling Project that analyzed gene expression in 240 DLBCL patient samples demonstrated reduced surface expression of MHC and CD58 molecules in DLBCL, which inhibits T-cell and NK-cell infiltration [50,51]. In addition, aberrant expression of anti-apoptotic and pro-apoptotic proteins in DLBCL reduces NK-cell cytotoxicity [52]. For instance, Bcl-2 protein is aberrantly overexpressed in 20% of DLBCL cases due to chromosomal translocation t(14;18) that juxtaposes Bcl-2 to the immunoglobulin heavy chain gene enhancers resulting in survival promotion [53]. Likewise, inactivating mutations in pro-apoptotic Noxa in DLBCL [54] contribute to decreased efficacy of NK-cell cytotoxicity [52]. Moreover, NHL cells express high levels of serpin B9, a serine protease inhibitor known to mitigate the tumoricidal activity of NK cells by blocking degranulation [55].

2.2. Studies on Acquired Resistance

2.2.1. ADCC-Resistance Models to Study Acquired Rituximab Resistance

The availability of in vitro models representing ADCC resistance in B-cell lymphoma is scarce. Czuczman et al. induced rituximab resistance (RR) in Raji (Burkitt lymphoma) and SU-DHL-4 (DLBCL) [56], whereas Jazirehi et al. in Daudi and Ramos (Burkitt lymphoma) cell lines [57] by incubations with escalating doses of RTX +/− human serum. Although these cell lines exhibit resistance to cell death mediated by RTX not only in CDC but also in ADCC mechanism, they were not generated by co-culture with NK cells and may not reflect the exact genetic and phenotypic changes that might have occurred upon this process. A study on RTX-resistant cell lines (RRCLs) demonstrated that point mutations in complementarity-determining regions (CDRs) of RTX induced enhanced ADCC and restored apoptosis in RR lymphoma cells as compared to wild-type RTX [58]. Therefore, enhancing the avidity of already existing mAbs might rescue or potentiate ADCC capability in r/r lymphoma.
Interesting data come from the study of Aldeghaither et al., who successfully induced ADCC resistance in squamous carcinoma cell line (A431) highly expressing epidermal growth factor receptor (EGFR) by co-culture with NK92-CD16V cells and incubation with anti-EGFR mAb–cetuximab [59]. This study proposed that ADCC resistance was primarily attributed to the impaired NK-cell activation caused by defects in immune synapse formation and cell-to-cell conjugation, rather than by a decrease in target antigen expression [59]. Similar observations were reported in acute lymphocytic leukemia (ALL) murine models, where the failure of treatment with T-cell receptor mimic monoclonal antibody (TCRm mAb) was not associated with target antigen loss or downregulation [60] but with a kinetic escape mechanism [61].

2.2.2. RTX-Mediated Changes in NK Cells

Rituximab causes phenotypical and functional alterations in NK cells from healthy donors [62]. Pre-incubation of non-malignant PBMCs with RTX led to NK-cell degranulation, downregulation of CD16, and depletion of healthy B cells, whereas in CD56dim cells, the expression of NK-cell activation markers, i.e., CD69 and CD137 was significantly increased, as compared to untreated cells. Another study demonstrated that CD16 downregulation is partially caused by its internalization upon interaction with RTX- or ofatumumab-coated B cells [63]. Defects in the cytotoxic potential of NK cells were found to correlate with some cross-inhibited NK-activating receptors, i.e., NKG2D, NKp46, 2B4, DNAM-1, which rendered primary NK cells from CLL patients hyporesponsive [63]. Further characterization of hyporesponsive NK cells revealed defects in their ability to phosphorylate some signaling molecules essential for NK-cell mediated cytotoxicity, i.e., SLP-76, PLCγ2, and Vav1 and consequently to prolonged recruitment of phosphatase SHP-1, which promoted desensitization of activated receptors [63].
Altogether, impaired basal ADCC activity might be associated with treatment failure in the future, whereas the negative impact of therapy-induced alterations of NK cells may partially explain the development of ADCC resistance.

2.3. Solutions to Improve ADCC Efficacy

2.3.1. Fc Receptor Engineering

Modification of Fc fragment of mAb to improve its binding affinity to FcγRIIIa proved to stimulate enhanced ADCC and CDC against lymphoma and leukemia cells as compared to unmodified RTX [64]. Fc-engineering strategy has been also employed in the generation of anti-CD20 mAb– obinutuzumab (OBI) currently registered for the treatment of CLL and FL [65]. OBI demonstrated markedly enhanced binding affinity to FcγRIIIa and dramatically improved potency of ADCC in comparison to RTX. Importantly, OBI-mediated activation of NK cells was reported to occur irrespective of KIR/HLA interactions [66,67]. Although OBI (Gazyva) combined with CHOP (G-CHOP) was not superior over R-CHOP in clinical trials involving previously untreated DLBCL patients (NCT01659099, NCT01287741), some benefit of G-CHOP in terms of PFS was associated with GCB subtype of DLBCL [68,69]. Nevertheless, the advantage of OBI over RTX has been reported in a large cohort of 1202 advanced-stage FL patients [70] and in 781 previously untreated CLL patients [71].
Another Fc-engineered anti-CD20 mAb–ocaratuzumab (AME-133v, currently in clinical trials–NCT01858181), showed more potent ADCC function compared to RTX in several clinical trials involving r/r FL patients [72,73], also in association with unfavorable FcγRIIIa-F genotype [74,75]. An improved binding affinity to FcγRIIIa and potentiation of NK cell-mediated ADCC was as well achieved by the introduction of S239D and I332E amino acid substitutions to generate anti-CD19 mAb–tafasitamab (MOR208) [76], which was highly effective in the treatment of r/r DLBCL and r/r FL patients [77].

2.3.2. Supplementation with Cytokines

Many efforts have been taken to augment NK-cell anti-tumor responses. Stimulation of NK cells with ALT-803, a super agonist of IL-15, a cytokine stimulating proliferation of these cells, resulted in enhanced ADCC against B-cell lymphoma in vitro and in vivo [78]. It has been demonstrated that supplementation with vitamin D3 significantly enhanced the efficacy of NK cell-mediated ADCC in elderly DLBCL patients and that rituximab-treated individuals suffering from vitamin D3 deficiency have worse survival [79]. Therefore, it has been suggested that vitamin D3 positively influences rituximab-mediated ADCC, and attempts to determine optimal serum levels of 25-OH-D3 for maximum NK-cell activity have been undertaken [80]. Unfortunately, a mechanism responsible for improving ADCC by vitamin D3 has not been published thus far.
Although molecular mechanisms underlying resistance to ADCC have not yet been thoroughly investigated specifically in DLBCL, available literature data suggest some factors involved. The existing ADCC in vitro and in vivo models propose that alterations in signaling pathways, defective NK-cell activation, and kinetics of cancer cell growth rather than target antigen loss might potentially play a crucial role in the development of resistance.

3. Resistance to Complement-Dependent Cytotoxicity

CDC is mediated by the classical pathway of the complement system upon binding of C1 complex to RTX-opsonized cells. This triggers activation of the complement cascade which results in the formation of the membrane attack complex (MAC) in the target cell membrane leading to a loss of membrane integrity and cell lysis [81].
Although ADCC is suggested as the main mechanism of action of RTX in clinics [15,16,17], the importance of complement is also postulated [82]. CDC seems to play a predominant role in mediating rituximab-induced cell death in CLL [83] due to high leukocytosis and therefore high accessibility of tumor cells to complement complex, which is reflected by the rapid exhaustion of complement and C3b(i) deposition [84]. In addition, side effects after infusion can be attributed to complement cascade activation [83]. Moreover, an extensive body of research demonstrated that CLL patients refractory to rituximab are often characterized with complement deficiencies, correlates with advanced disease, and shorter survival time [85,86,87]. In line with these findings, administration of individual complement components restores CDC efficacy in complement-deficient patients [88], and concurrent administration of rituximab and fresh frozen plasma remarkably enhanced the therapeutic activity of RTX [89]. In addition, CLL cells resistant to rituximab have significantly higher expression of complement regulatory proteins (CRPs) than the sensitive cells [90]. In CLL, primary resistance to CDC is mediated also by the presence of soluble protective protein complement factor H (CFH) that prevents the formation and deposition of additional C3b and propagation of the complement cascade. Additionally, inhibiting or inactivating CFH with monoclonal antibody permits activation of complement leading to cell lysis and death in resistant CLL cells [91].
In DLBCL patients, given the scarce presence of malignant cells in the circulation and therefore lower accessibility of tumor cells to complement complex, it would seem that the main benefit of incorporating rituximab into CHOP relies on its ability to chemosensitize malignant cells. However, it has been demonstrated in a disseminated lymphoma tumor mouse model that C1q-deficient mice exhibited a defective response to RTX, while depletion of either NK cells or granulocytes did not hamper rituximab efficacy [92]. In a human-CD20 transgenic mouse model, Gong et al. showed that circulating B cells were depleted through ADCC, while B cells in the marginal zone of lymph nodes were lysed primarily by CDC [93].
Some studies show an essential role of complement in determining the clinical efficacy of rituximab also in DLBCL, as high expression of CD59 correlates with poor OS and PFS [94]. CD59 is one of CRPs that blocks the formation of MAC, which is the last step of complement activation and thus prevents cell lysis [95]. In line with these findings, the levels of miR-224, a negative regulator of CD59 expression, are significantly lower in DLBCL patients with poor OS and PFS [94].
The engagement of CDC in the clinical success of rituximab in NHL is a matter of controversy, as it has been suggested that complement activation can be detrimental to antibody immunotherapy in murine lymphoma models. It has been demonstrated that breakdown products of the complement component C3 block rituximab activity and render NK cells less active in ADCC [96]. In line with that, in vitro studies demonstrated that NK-cell activation and RTX-induced ADCC are inhibited by the C3b component of complement [97]. Moreover, polymorphisms in C1q gene allele were recently shown as important factors that possibly influence the success of therapy with rituximab in NHL. It has been demonstrated that FL patients homozygous for the C1qA[276]G allele, which most probably correlates with higher serum levels of C1q and thus higher complement activity, have a significantly shorter time to progression than A carriers [98]. An analysis of DLBCL patients demonstrated that homozygotes for the C1qA[276]G allele had not only shorter PFS but also decreased OS and CR rates [99]. Nevertheless, the role of complement in determining the efficacy of rituximab in NHL remains an open question and requires further investigation.

3.1. The Role of the Cell Membrane Composition in Determining Primary Resistance to Rituximab

The binding of rituximab to CD20 induces the translocation of the antibody–antigen complex to the lipid rafts [100], highly dynamic domains in the cell membrane rich in cholesterol and sphingolipids that serve as docking points for signal transduction [101]. It is recognized that the composition of the lipid raft domains may play an important role in determining resistance to rituximab [102]. In CLL patients, resistance to rituximab was shown to be associated with aberrant recruitment of Csk-binding protein, leading to a lack of sphingomyelin and ganglioside M1 at the outer leaflet of the plasma membrane of tumor cells [103]. Furthermore, low ganglioside M1 levels in CLL and MCL patients resulted in a lack of sensitivity to RTX-induced CDC in vitro [104]. Moreover, a study on primary CLL samples demonstrated that resistance to CDC is strongly associated with cell surface α2-6 hypersialylation induced by increased activity of α2-6 sialyl transferase and that the removal of terminal sialic acids enhances the activity of rituximab-induced CDC [83]. Importantly, hypersialylation that leads to binding of factor H is a known factor protecting erythrocytes from complement-induced lysis [105]. Observations from MCL primary samples and cell lines show that malignant B cells produce aberrant lipid rafts leading to impaired signaling transduction as a way to avoid apoptosis [106]. Similar findings have been published for DLBCL, where the altered organization of lipid rafts led to the resistance to conventional chemotherapy [107]. Additionally, the presence of plasma membrane cholesterol was found to promote conformational changes in CD20 antigen, which are critical for the effective binding of anti-CD20 mAbs and thus potent RTX-mediated CDC [108]. Therefore, it is highly probable that membrane composition influences response to rituximab also in this type of lymphoma.

3.2. Rituximab-Resistant Cell Lines as a Tool to Study the Mechanisms of Acquired Resistance to CDC

Already 10 years after rituximab introduction to the clinic, two large studies on r/r DLBCL patients clearly demonstrated that rituximab used as the first-line treatment leads to substantial alterations in the biology of malignant cells and makes them resistant to subsequent salvage regimens [109,110]. Early relapses following rituximab-containing first-line therapy have a poor prognosis [109] and prior exposure to rituximab was demonstrated to be an independent adverse prognostic factor for both PFS and OS in r/r DLBCL [110].
To better understand the mechanisms of resistance, cell line models have been generated by Czuczman et al. [56], Barth et al. [111], Takei et al. [112], and Jazirehi et al. [57]. They were generated by repeated exposure to elevated concentrations of rituximab or rituximab plus human serum as a source of complement, thus they mimic the development of resistance to CDC-dependent cell death. The generated cell lines have become important models for studying the molecular mechanisms of resistance induced by prolonged exposition to rituximab.
First of all, they enabled to determine that CD20 downregulation occurs following exposure to rituximab is regulated on both transcriptional and posttranslational levels [56,113]. Second, upregulation of complement inhibitory proteins (CD46, CD55, CD59) in RRCLs has been reported [56]. This corresponds with the data obtained from in vitro functional analyses on primary malignant cells from both NHL [114] and CLL [90], where the levels of CRPs (mainly CD55) were suggested as markers for clinical response to rituximab treatment.

3.3. CD20 Downregulation in Acquired Resistance to Rituximab

Resistance to rituximab and R-CHOP results as a consequence of molecular alterations of both target malignant cells, effector immune cells, and the influence of tumor microenvironment. One of the most immediate causes of resistance to anti-CD20 mAbs are decreased CD20 levels [13,115]. Yet, the importance of CD20 downregulation in real-world settings has not been sufficiently documented and it is based on case reports [84,115,116,117,118,119]. Even if the response rate at relapse to rituximab in prior responders is below 50%, these patients are not routinely re-biopsied [117].
CD20 levels have been shown by several groups to correlate with the efficacy of rituximab (reviewed in [13]). Experimental in vitro studies using cell clones with divergent CD20 expression demonstrated a direct correlation between the number of CD20 molecules on the cell surface and CDC efficacy [120]. Intriguingly, no such correlation has been observed in ADCC efficacy [120]. However, our observations on increasing CD20 levels with small-molecule inhibitors showed improved efficacy of rituximab in SCID mice, where NK-cell mediated ADCC is one of the leading effector mechanisms [121].
The biological function of CD20 is still unclear despite years of extensive research. CD20 is described as a component of a store-operated calcium entry pathway activated by B-cell receptor (BCR) [122]. Until now, the existing body of evidence has suggested that CD20 loss in humans and in murine models leads to a relatively mild phenotype (reviewed in [13]) with no alterations in B-cell survival nor in proliferation rate [123]. Therefore, as CD20 is not indispensable for cell survival, it is not surprising that tumor cells downregulate this target molecule. The mechanisms leading to CD20 downregulation has been reported by several groups based both on clinical observations and preclinical in vitro studies (reviewed in [13]) and include all steps of protein synthesis–transcription [124,125,126], translation as well as post-transcriptional and post-translational modifications [121] and transport of CD20 to the cell membrane [108,127,128]. Moreover, the presence of a soluble form of CD20 protein reported in CLL patients [129] and CD20 secreted in exosomes in patients suffering from aggressive NHLs [130], which compete with CD20 for binding to rituximab, contributes to the reduction of the effectiveness of mAbs. There are also reports on trogocytosis—“shaving” and endocytosis of the antigen by phagocytic cells [41], but it has to be kept in mind that this mechanism has been proven only in in vitro settings.
The mechanisms of CD20 regulation and potential compounds that may be combined with rituximab in order to increase its efficacy by inducing CD20 upregulation are elegantly reviewed in [13,131]. A genome-wide study by de Jong et al. provides data on rituximab compatible drug-target combinations for DLBCL [132].

3.4. Hexamerization-Inducing Mutations as a Way to Improve CDC Efficacy

Recently, it has been demonstrated that Fc fragments of IgG antibodies form ordered Ab-hexamers following antigen binding [133]. The introduction of hexamerization-inducing mutations (E345 or E430) in IgG1 gene strongly augmented the ability of some Abs to induce CDC and ADCC against lymphoma cell lines and primary CLL samples [134]. Since both mutations were found to play a stimulatory role in inducing CDC, the presence of the conserved E residue might be responsible for the suppression of excessive complement activation [134].

4. Mechanisms of Chemoresistance following Rituximab-Treatment

Given the mentioned poor prognosis following the first-line treatment with rituximab in relapsed patients, the mechanisms of chemoresistance have also been thoroughly studied in the models of RRCLs. All the RRCLs are resistant to cisplatin, doxorubicin, paclitaxel, and vincristine due to decreased expression of the Bcl-2 family proteins, i.e., Bax, Bak, Bcl-2, and increased expression of Bcl-xL and Mcl-1 [135,136]. Therefore, it can be concluded that while a limited exposure of lymphoma cells to rituximab may lead to a rise in the number of pro-apoptotic proteins sensitizing malignant cells to chemotherapy, repeated exposure leads to the induction of an apoptosis-resistant phenotype [135]. Prolonged exposure of cell lines to rituximab leads also to downregulation of ICAM-1, which impairs cellular adhesion and aggregation [137]. Importantly, ICAM-1 was demonstrated to be a marker of positive response in NHL rituximab-treated patients. It has been postulated that ICAM-1 downregulation in RRCLs may contribute to the development of chemoresistance, as cell clustering upon rituximab-treatment increases the efficacy of chemotherapeutics [137].
RRCLs are also characterized by profound metabolic changes. It has been demonstrated that the acquirement of rituximab resistance is associated with deregulation in glucose metabolism caused by overexpression of hexokinase II (HKII), which performs dual functions of promoting glycolysis as well as inhibiting mitochondrial-mediated apoptosis [138]. These data have been next confirmed in gene expression profiling datasets from R-CHOP treated patients and higher levels of HKII are considered to correlate with poor PFS and OS in DLBCL [138].

5. The Role of Tumor Microenvironment in Promoting Resistance to Rituximab-Containing Regimens

Recently, more attention has been given to the role of the tumor environment in promoting tumor progression and influencing the clinical response to treatment [139]. In NHL patients where only a minor part of malignant cells is found in the circulation, the role of the microenvironment is particularly important [140]. For instance, mesenchymal stromal cells (MSCs) decrease the efficacy of anti-CD20 treatment by decreasing CD20 expression [141]. Moreover, adhesion to stromal cells protects malignant B cells from apoptosis induced by chemotherapeutics as well as from rituximab-induced apoptosis [142]. Blocking the interaction between tumor and stromal cells with natalizumab, which targets VLA-4 (integrin alfa-4-beta-1/CD49d), partially overcomes this protection [142].
Although bone marrow infiltration is not common in DLBCL, the studies of this milieu in the context of therapy efficacy need to be developed as bone marrow-derived MSCs have been shown to promote the growth and primary resistance of DLBCL cell lines to rituximab by secreting IL-6 and elevating IL-17A levels in in vitro settings [143]. Increased levels of IL-6 have been demonstrated in the immunohistochemistry of tumor tissues from DLBCL patients [143]. Importantly, rituximab itself may induce tumor cells to produce IL-6 and consequently induce T-regulatory cells (Tregs) to secrete IL-17, which leads to tumor growth and rituximab resistance in DLBCL patients [144]. Therefore, it has been postulated that rituximab could be a “double-edged sword” and induce resistance and poor prognosis in relapsed DLBCL patients [144].
Conversely, rituximab has been shown to promote long-time response due to its vaccinal effect as demonstrated in EL4-huCD20 murine models of lymphoma that were resistant to tumor expansion at rechallenge with RTX [145]. Further studies have shown that rituximab administration reverses the expansion of immunosuppressive Tregs and increases the number of pro-inflammatory T-helper 1 (Th1) cells. Rituximab treatment led to an increase in IL-12, increased numbers of activated myeloid dendritic cells, and CD4-positive effector memory T cells in treated animals [146].
As DLBCL cells predominantly reside in lymph nodes, the stromal factors implicated in the organization of extracellular matrix (ECM) and the cells of innate immunity, especially macrophages, dendritic cells (DCs), and NK cells are involved in the physiopathology and drug response of DLBCL [147]. A study by Lenz et al. described a “stromal” signature characterized by high deposition of ECM proteins, e.g., fibronectin, secrete protein acid rich in cysteine (SPARC), and vast infiltration of myelo-monocytic cells [148]. Such signature leads to prolonged survival following R-CHOP suggesting an adjuvant role of macrophages in promoting rituximab cytotoxicity. In contrast, the presence of an immunosuppressive population of M2-like macrophages was shown to correlate with shorter survival in R-CHOP-treated cohorts [149]. Therefore, therapeutic strategies aiming at promoting the differentiation to M1 macrophages may improve prognosis in DLBCL.

Targeting Immune Microenvironment as a Way to Improve Rituximab Efficacy

In light of the above, modifying the tumor environment with immunomodulatory drugs seems a rational solution. One of such drugs is lenalidomide, currently approved by FDA for MCL treatment [150]. In an international phase II trial of r/r B-NHL patients, lenalidomide as a single agent had ORR of 35% [151]. Combination with rituximab seems to be synergistic and highly effective, as 90% of previously untreated advanced-stage low-grade NHL patients reported a CR [152]. This effect may be due to lenalidomide-mediated activation of DCs and increased recruitment of NK to the tumor site [153]. Multiple phase I/II trials testing combinations of lenalidomide with R-CHOP in hope of improving long-term outcomes are underway.

6. Other Treatment Modalities as Possible Solutions for Relapsed/Refractory Patients

6.1. Novel Targets for mAbs

Despite the success of rituximab as a safe and efficient therapeutic option, the incidence of resistance encouraged searching for other molecular targets for immunotherapies in the management of DLBCL (Figure 2). These novel targets, including CD19, CD22, CD40, CD80, and CD79b, have been comprehensively reviewed in [7,154,155,156]. The most successful mAbs in recent times are the ones targeting CD19–tafasitamab registered recently in a combination with lenalidomide and antibody–drug conjugates (ADCs), i.e., loncastuximab tesirisine-lpyl and an anti-CD79b ADC–polatuzumab vedotin. The approval of polatuzumab vedotin in combination with bendamustine and rituximab in r/r DLBCL was preceded by a clinical trial that for the first time after several years demonstrated the benefit in OS to an experimental arm of the study [154]. Besides the above-mentioned targets, an attractive alternative is the use of CD37-directed mAbs, which are currently in clinical trials and we have recently reviewed them in [157].

6.2. Bispecific Antibodies

A way to improve the efficacy of treatment response is to employ dual antigen targeting with bispecific Abs (bsAbs) to stimulate a more potent anti-tumor response, increase binding specificity and overcome the low efficacy of single-epitope targeting. Moreover, the inclusion of single-chain variable fragment (scFv) of an anti-CD16 Ab into bsAb constructs has been employed to enhance CD16-positive effector cell recruitment [158]. In addition, utilizing the cytotoxic potential of T cells in synergy with conventional mechanisms of mAb-mediated cancer cell death has been recently exploited in the generation of bispecific T-cell engagers (BiTEs). Examples of bsAbs and BiTEs, their mode of action, and efficacy are summarized in Table 1.

6.3. Antibody–Drug Conjugates

Antibody–drug conjugates are composed of specific mAbs linked to cytotoxic agents, the so-called free payload, that exert their cytotoxic action upon internalization [171]. Since the cytotoxic agents are released at the tumor site the likelihood of systemic exposure and toxicity is therefore significantly lower. Additional advantage comes from the immunogenic cell death, as the release of damage-associated molecular patterns (DAMPs) by the tumor cells is known to stimulate antigen-presenting cells (APCs) [172].
Several specific markers of B cells that may be used as targets for ADCs have been identified in in vitro studies, i.e., CD19, CD20, CD21, CD22, CD72, CD79b, and CD180 [173]. By now, a few ADCs have been registered as treatment options for B-cell malignancies, i.e., brentuximab vedotin (CD30-positive Hodgkin lymphoma (HL); anaplastic large-cell lymphoma (ALCL) [174], inotuzumab ozogamicin (r/r B-cell precursor ALL) [175], polatuzumab vedotin (r/r DLBCL) [176] and gemtuzumab ozogamicin (for CD33-positive acute myeloid leukemia (AML)). ADCs against various lymphoma types with the highest clinical potential are summarized in Table 2.
Based on in vitro studies and xenograft models of repeated exposure to consecutively increasing concentrations of ADCs, several leading cellular mechanisms, by which the acquired resistance to ADCs occurs were described, each elegantly reviewed by Loganzo et al. [177]. The first resistance mechanism is antigen downregulation or mutation [178]. Moreover, the development of resistance was found to be correlated with the overexpression of ATP-binding cassette (ABC) transporters such as MDR1, MRP1, and BCRP, which play a significant role in the cellular efflux of multiple antineoplastic agents [179] such as MMAE, DM1, and ozogamicin [178]. Brentuximab vedotin-refractory HL, patients were found to have upregulation of the multidrug resistance MDR1 gene [180]. The expression of ABC transporters is increased also in DLBCL cells, where it correlates with a worse response to treatment [181].
Several attempts are undertaken to enhance the therapeutic potential of ADCs [182]. Alterations within the antibody (to obtain sufficient antigen affinity and specificity), linker or the payload (DNA-interactive payload classes or microtubule inhibitors, e.g., calicheamicins, auristatins, and maytansinoids) increase ADC efficacy [182,183]. Another approach is including ADCs in multi-drug schemes, e.g., testing inotuzumab ozogamicin with R-CHOP for the treatment of DLBCL patients unfit for anthracycline has been evaluated in NCT01679119 study [184].
Table 2. Antibody–drug conjugates in clinical trials.
Table 2. Antibody–drug conjugates in clinical trials.
NameTarget AntigenCytotoxic Payload and its Mechanism of ActionPossible Application of ADCClinical Efficacy of ADC
Polatuzumab vedotin (PV)CD79bMonomethyl auristatin E. (MMAE) with a cleavable linker for disruption of microtubule networkDLBCL–registrationPV was combined with RTX and bendamustine (pola-BR) and compared to bendamustine and RTX treatment alone. Pola-BR patients had a significantly higher CR rate (40.0% vs. 17.5%; p = 0.026) and longer PFS [176].
Brentuximab vedotin (BV)CD30auristatin E. (MMAE) with a cleavable linker for disruption of microtubule networkHL–registration ALCL-registration75% and 86% OR rates in r/r HL and ALCL, consecutively [185].
Inotuzumab ozogamicin
(CMC-544)
CD22Calicheamicin with a cleavable (acid-labile) linker for disruption of double-stranded DNA in the nucleusNHL
B-ALL–registration
39% ORR in r/r/B-ALL (NCT01134575)
CMC-544 in combination with RTX resulted in longer survival in a disseminated B-cell lymphoma murine model [186].
AGS67ECD37Monomethyl auristatin E. (MMAE) with protease-cleavable linker for disruption of microtubule networkNHL
CLL
Until now only one clinical trial was completed, the safety of AGS67E was demonstrated in patients with r/r lymphoid malignancies (NCT02175433) [187].
Pinatuzumab vedotinCD22Monomethyl auristatin E. (MMAE) with a cleavable linker (with sulfhydryl groups) for disruption of microtubule networkNHLObjective responses were observed in DLBCL (9/25); CR in 2/8 patients treated with pinatuzumab vedotin and RTX [188].
Naratuximab emtansine
(IMGN529)
CD37Emtansine (DM1) with a non-cleavable linker for disruption of microtubule polymerizationr/r B-cell lymphomasIn a phase 1 study on B -NHL, a reduction in lymphocyte count was observed after the second dosing of IMGN529 [189].
Denintuzumab mafodotin
(SGN-CD19A)
CD19Monomethyl auristatin F with a cleavable linker for disruption of tubulin polymerizationNHLIn a phase 1 study SGN-CD19A exerted durable responses in heavily pre-treated NHL patients; 56% objective responses were achieved for relapsed patients with a CR rate of 40% [190].
Coltuximab ravtansine
(SAR3419)
CD19Ravtansine (DM4) with a cleavable linker (with disulfide groups) for disruption of tubulin polymerizationDLBCLIn a phase 2 study on r/r/DLBCL, ORR was achieved in 43% of patients (18/41), where PFS and OS were 4.4 and 9.2 months, respectively [191].

6.4. CAR-T Cells

In recent years, considerable progress in hematooncology has been achieved by developing adoptive therapies based on T cells modified with chimeric antigen receptors (CARs) [192,193]. CARs targeting CD19 have been a breakthrough in the treatment of ALL and DLBCL. The results of clinical trials have been especially promising in ALL, where high CR (up to approx. 80%) was observed, while it was much lower (approx. 50%) in r/r DLBCL patients. Still, it has to be underlined that in r/r DLBCL prognosis is dismal; thus, each therapeutic option counts. The encouraging results of the clinical trials resulted in the registration of four different CAR CD19 formulations varying in the composition of the intracellular co-stimulation domain, three of which, namely: tisagenlecleucel, axicabtagene ciloleucel, and lisocabtagene maraleucel are approved for the treatment of DLBCL patients [193,194,195].
However, almost half of patients with high-grade lymphomas progress after CD19 CAR-T cell therapy and require additional treatment [196]. The mechanisms of tumor escape are related to host factors affecting CAR proliferation and persistence, features of CAR-T product and their toxicity, and tumor intrinsic factors, such as loss of target epitope expression or trogocytosis [192]. By now, the information on the mechanisms of resistance to CAR-based therapy comes mainly from observations in ALL.
In ALL, antigen loss is one of the most frequent reasons for relapse after CD19 CAR-T therapy. Disease relapse can occur by 12 months after infusion of CAR-T cells in up to 50% of patients [197]. Relapse connected with CD19 antigen loss is also observed in the case of NHLs but is less frequent than in ALL [198].
Thus far, two main mechanisms of antigen loss were described in ALL–antigen escape and lineage switch. In the first case, after remission in response to CD19 CAR-T therapy, patients lacking surface CD19 antigen relapse but with a phenotypically similar disease. In contrast, lineage switch describes the opposite situation, when relapse is a phenotypically different malignancy, but genetically similar [199]. That occurs because of several splicing variants of CD19 lacking the CAR-recognizing epitope [200,201]. It was demonstrated that relapsed ALL is characterized by a mutation in exon 2 of CD19 and the expression of other CD19 isoforms. This resistance mechanism supports BCP-ALL proliferation and prevents the killing by CD19 CAR-T [202]. Recently, the role of hypermethylation of CD19 promoter in antigen loss has been demonstrated in an in vivo CD19-negative recurrence model of CLL following CD19 CAR-T therapy [203]. Less frequently, the tumor escape is mediated by low antigen density or trogocytosis [192,204]. Moreover, as the effectors acquire CD19 expression at their surface, they are in consequence eliminated by fratricide killing [204]. Nevertheless, the recent study by Jain et al. has questioned the role of CD19 downregulation/loss in determining resistance to CD19 CAR-T in DLBCL and underlined the role of immune exhaustion [205]. By characterizing the genomic signature of DLBCL primary samples obtained from patients that subsequently underwent CD19 CAR-T administration, they discovered some tumor-intrinsic genomic alterations such as chromothripsis (aberrantly reassembled chromosomes leading to aneuploidies), the mutational activity of APOBEC, deletions of RB1 or RHOA that according to the authors correlate with exhausted immune microenvironments and resistance/poor response to CAR therapy [205].
CD19 loss after CAR-T therapy in ALL can be accompanied by the downregulated expression of some surface antigens, such as CD20 and CD22 [206]. A similar effect of CD22 downregulation with CD19 loss after CD19 CAR-T treatment was observed in DLBCL patients [207]. Moreover, the CD58 aberration (lack of expression or mutation) is observed in some r/r DLBCL patients after CD19 CAR-T treatment. CD58, as a ligand for CD2 on T cells that provides the costimulatory signal, has been linked to resistance to CD19 CAR-T and progression after this therapy in DLBCL [208]. There is also some evidence that 10–20% of patients with ALL have primary resistance to CD19 CAR-T. This type of resistance can be correlated with lower expression of death receptors connected with apoptosis pathways in neoplastic cells as demonstrated in in vitro model testing CRISPR-based genome-wide loss-of-function screen in Nalm-6 ALL cell line [209]. The potential strategies against resistance to CAR T-cell therapy in hematological malignancies, including modifications in the structure of CARs, novel therapeutic targets, and safety-related issues have been elegantly reviewed in [192,193,210].

7. Perspectives

Hematological malignancies derive from immune cells and are surrounded by other components of the immune system. This provides many opportunities to interact with the tumor environment, locally transported antibodies, and the complement system. These characteristics shape unique conditions for the use of immunotherapy, which in the last two decades has become the mainstream for many types of hematological malignancies, including the most frequent one—DLBCL.
Despite extensive research on novel therapeutic options, R-CHOP combination remains the core of therapy in DLBCL. While the frequency of primary resistance to R-CHOP is relatively low (approx. 10%), secondary resistance acquired in the course of therapy leads to relapse in around one-third of patients [211]. The situation is particularly delicate in the case of r/r DLBCL patients, as their prognosis is dismal. In this population, the overall survival amounts to 6 months [5]. Therefore, r/r DLBCL patients urgently need novel therapeutic options. Given the genetic heterogeneity of this malignancy, targeting common B-specific antigens present on the cell surface with monoclonal antibodies or CAR-modified T- and NK-cells seems a rational option. Nevertheless, the years of employment of rituximab as a core of first-line treatment and recent observations on the mechanisms of resistance to CAR-T cell therapy suggest that treatment resistance is an inevitable consequence of applying immunotherapeutics. Therefore, a question arises: which therapeutic schemes will cause maximal efficacy against malignant cells in r/r patients? A thorough study of the mechanisms of primary and acquired resistance to immunotherapy may lead to optimization of the already existing compounds as demonstrated by the introduction of hexamerization-inducing mutations sensitizing to CDC and Fc-engineering of mAbs, leading to increased ADCC. Analyses of resistant cellular models and primary material from r/r patients can reveal immunotherapy-induced changes in the phenotype of malignant cells, which can be further addressed by designing specific mAbs and CAR-modified effector cells.

Author Contributions

Conceptualization, A.K. and M.B.; methodology, A.K. and M.B.; investigation, A.K., M.B., M.K. (Marta Krawczyk), M.W. and M.K. (Matylda Kubacz); writing—original draft preparation A.K., M.B., M.K. (Marta Krawczyk), M.W. and M.K. (Matylda Kubacz); writing—review and editing, A.K., M.B., M.K. (Marta Krawczyk), M.W., M.K. (Matylda Kubacz) and A.S.; visualization, A.S. and A.K.; supervision, M.W.; funding acquisition, M.B and M.W. All authors have read and agreed to the published version of the manuscript.

Funding

This project was supported by European Research Council 805038/STIMUNO/ERC-2018-STG (M.W.), Polish National Science Centre 2019/35/D/NZ5/01191 (M.B.) and 2020/39/O/NZ6/01434 (M.W. and M.K. (Marta Krawczyk)), Medical University of Warsaw 1M19/1/M/MG/N/21/21 (M.K. (Matylda Kubacz)). The Ministry of Science and Higher Education within “Regional Initiative of Excellence” program 013/RID/2018/19, project budget 12 000 000 PLN (M.W.).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Thandra, K.C.; Barsouk, A.; Saginala, K.; Padala, S.A.; Barsouk, A.; Rawla, P. Epidemiology of Non-Hodgkin’s Lymphoma. Med. Sci. 2021, 9, 5. [Google Scholar] [CrossRef] [PubMed]
  2. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef] [PubMed]
  3. Shaw, J.; Harvey, C.; Richards, C.; Kim, C. Temporal Trends in Treatment and Survival of Older Adult Diffuse Large B-Cell Lymphoma Patients in the SEER-Medicare Linked Database. Leuk. Lymphoma 2019, 60, 3235–3243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Rovira, J.; Valera, A.; Colomo, L.; Setoain, X.; Rodríguez, S.; Martínez-Trillos, A.; Giné, E.; Dlouhy, I.; Magnano, L.; Gaya, A.; et al. Prognosis of Patients with Diffuse Large B Cell Lymphoma Not Reaching Complete Response or Relapsing after Frontline Chemotherapy or Immunochemotherapy. Ann. Hematol. 2015, 94, 803–812. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Crump, M.; Neelapu, S.S.; Farooq, U.; Van Den Neste, E.; Kuruvilla, J.; Westin, J.; Link, B.K.; Hay, A.; Cerhan, J.R.; Zhu, L.; et al. Outcomes in Refractory Diffuse Large B-Cell Lymphoma: Results from the International SCHOLAR-1 Study. Blood 2017, 130, 1800–1808. [Google Scholar] [CrossRef] [PubMed]
  6. Nowakowski, G.S. Recently Approved Drugs Herald a New Era in Therapy for Diffuse Large B-Cell Lymphoma. Clin. Adv. Hematol. Oncol. 2021, 19, 284–287. [Google Scholar]
  7. Wang, L.; Li, L.; Young, K.H. New Agents and Regimens for Diffuse Large B Cell Lymphoma. J. Hematol. Oncol. 2020, 13, 175. [Google Scholar] [CrossRef]
  8. He, M.Y.; Kridel, R. Treatment Resistance in Diffuse Large B-Cell Lymphoma. Leukemia 2021, 35, 2151–2165. [Google Scholar] [CrossRef]
  9. Engelhard, M. Anti-CD20 Antibody Treatment of Non-Hodgkin Lymphomas. Clin. Immunol. 2016, 172, 101–104. [Google Scholar] [CrossRef]
  10. Salles, G.; Barrett, M.; Foà, R.; Maurer, J.; O’Brien, S.; Valente, N.; Wenger, M.; Maloney, D.G. Rituximab in B-Cell Hematologic Malignancies: A Review of 20 Years of Clinical Experience. Adv. Ther. 2017, 34, 2232–2273. [Google Scholar] [CrossRef] [Green Version]
  11. Winiarska, M.; Glodkowska-Mrowka, E.; Bil, J.; Golab, J. Molecular Mechanisms of the Antitumor Effects of Anti-CD20 Antibodies. Front. Biosci. (Landmark Ed) 2011, 16, 277–306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Johnson, P.; Glennie, M. The Mechanisms of Action of Rituximab in the Elimination of Tumor Cells. Semin. Oncol. 2003, 30, 3–8. [Google Scholar] [CrossRef] [PubMed]
  13. Pavlasova, G.; Mraz, M. The Regulation and Function of CD20: An “Enigma” of B-Cell Biology and Targeted Therapy. Haematologica 2020, 105, 1494–1506. [Google Scholar] [CrossRef] [PubMed]
  14. Alas, S.; Emmanouilides, C.; Bonavida, B. Inhibition of Interleukin 10 by Rituximab Results in Down-Regulation of Bcl-2 and Sensitization of B-Cell Non-Hodgkin’s Lymphoma to Apoptosis. Clin. Cancer Res. 2001, 7, 709–723. [Google Scholar]
  15. Jazirehi, A.R.; Bonavida, B. Cellular and Molecular Signal Transduction Pathways Modulated by Rituximab (Rituxan, Anti-CD20 MAb) in Non-Hodgkin’s Lymphoma: Implications in Chemosensitization and Therapeutic Intervention. Oncogene 2005, 24, 2121–2143. [Google Scholar] [CrossRef] [Green Version]
  16. Vega, M.I.; Baritaki, S.; Huerta-Yepez, S.; Martinez-Paniagua, M.A.; Bonavida, B. A Potential Mechanism of Rituximab-Induced Inhibition of Tumor Growth through Its Sensitization to Tumor Necrosis Factor-Related Apoptosis-Inducing Ligand-Expressing Host Cytotoxic Cells. Leuk. Lymphoma 2011, 52, 108–121. [Google Scholar] [CrossRef]
  17. Clynes, R.A.; Towers, T.L.; Presta, L.G.; Ravetch, J.V. Inhibitory Fc Receptors Modulate in Vivo Cytoxicity against Tumor Targets. Nat. Med. 2000, 6, 443–446. [Google Scholar] [CrossRef]
  18. Hamaguchi, Y.; Xiu, Y.; Komura, K.; Nimmerjahn, F.; Tedder, T.F. Antibody Isotype-Specific Engagement of Fcγ Receptors Regulates B Lymphocyte Depletion during CD20 Immunotherapy. J. Exp. Med. 2006, 203, 743–753. [Google Scholar] [CrossRef] [Green Version]
  19. Umana, P.; Moessner, E.; Bruenker, P.; Unsin, G.; Puentener, U.; Suter, T.; Grau, R.; Schmidt, C.; Gerdes, C.; Nopora, A.; et al. Novel 3rd Generation Humanized Type II CD20 Antibody with Glycoengineered Fc and Modified Elbow Hinge for Enhanced ADCC and Superior Apoptosis Induction. Blood 2006, 108, 229. [Google Scholar] [CrossRef]
  20. Boross, P.; Jansen, J.H.M.; de Haij, S.; Beurskens, F.J.; van der Poel, C.E.; Bevaart, L.; Nederend, M.; Golay, J.; van de Winkel, J.G.J.; Parren, P.W.H.I.; et al. The in Vivo Mechanism of Action of CD20 Monoclonal Antibodies Depends on Local Tumor Burden. Haematologica 2011, 96, 1822–1830. [Google Scholar] [CrossRef]
  21. Vivier, E.; Tomasello, E.; Baratin, M.; Walzer, T.; Ugolini, S. Functions of Natural Killer Cells. Nat. Immunol. 2008, 9, 503–510. [Google Scholar] [CrossRef] [PubMed]
  22. Cooper, M.A.; Fehniger, T.A.; Caligiuri, M.A. The Biology of Human Natural Killer-Cell Subsets. Trends Immunol. 2001, 22, 633–640. [Google Scholar] [CrossRef]
  23. Alderson, K.L.; Sondel, P.M. Clinical Cancer Therapy by NK Cells via Antibody-Dependent Cell-Mediated Cytotoxicity. J. Biomed. Biotechnol. 2011, 2011, 379123. [Google Scholar] [CrossRef]
  24. Wang, W.; Erbe, A.K.; Hank, J.A.; Morris, Z.S.; Sondel, P.M. NK Cell-Mediated Antibody-Dependent Cellular Cytotoxicity in Cancer Immunotherapy. Front. Immunol. 2015, 6, 368. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Li, Y.; Huang, K.; Liu, L.; Qu, Y.; Huang, Y.; Wu, Y.; Wei, J. Effects of Complement and Serum IgG on Rituximab-Dependent Natural Killer Cell-Mediated Cytotoxicity against Raji Cells. Oncol. Lett. 2019, 17, 339–347. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Koene, H.R.; Kleijer, M.; Algra, J.; Roos, D.; von dem Borne, A.E.G.K.; de Haas, M. FcγRIIIa-158V/F Polymorphism Influences the Binding of IgG by Natural Killer Cell FcγRIIIa, Independently of the FcγRIIIa-48L/R/H Phenotype. Blood 1997, 90, 1109–1114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Cartron, G.; Dacheux, L.; Salles, G.; Solal-Celigny, P.; Bardos, P.; Colombat, P.; Watier, H. Therapeutic Activity of Humanized Anti-CD20 Monoclonal Antibody and Polymorphism in IgG Fc Receptor FcγRIIIa Gene. Blood 2002, 99, 754–758. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Weng, W.-K.; Levy, R. Two Immunoglobulin G Fragment C Receptor Polymorphisms Independently Predict Response to Rituximab in Patients with Follicular Lymphoma. J. Clin. Oncol. 2003, 21, 3940–3947. [Google Scholar] [CrossRef]
  29. Kim, D.H.; Jung, H.D.; Kim, J.G.; Lee, J.-J.; Yang, D.-H.; Park, Y.H.; Do, Y.R.; Shin, H.J.; Kim, M.K.; Hyun, M.S.; et al. FCGR3A Gene Polymorphisms May Correlate with Response to Frontline R-CHOP Therapy for Diffuse Large B-Cell Lymphoma. Blood 2006, 108, 2720–2725. [Google Scholar] [CrossRef] [Green Version]
  30. Ahlgrimm, M.; Pfreundschuh, M.; Kreuz, M.; Regitz, E.; Preuss, K.-D.; Bittenbring, J. The Impact of Fc-γ Receptor Polymorphisms in Elderly Patients with Diffuse Large B-Cell Lymphoma Treated with CHOP with or without Rituximab. Blood 2011, 118, 4657–4662. [Google Scholar] [CrossRef]
  31. Büyükkurt, N.; Özcan, M.A.; Ergene, Ü.; Payzın, B.; Tunalı, S.; Demirkan, F.; Özsan, H.; Pişkin, Ö.; Ündar, B. The Effect of FcγRIIIA Gene Polymorphism on the Treatment of Diffuse Large B-Cell Non-Hodgkin Lymphoma: A Multicenter Prospective Observational Study. Turk. J. Haematol. 2015, 32, 152–157. [Google Scholar] [CrossRef] [PubMed]
  32. Liu, F.; Ding, H.; Jin, X.; Ding, N.; Deng, L.; He, Y.; Zhu, J.; Song, Y. FCGR3A 158V/F Polymorphism and Response to Frontline R-CHOP Therapy in Diffuse Large B-Cell Lymphoma. DNA Cell Biol. 2014, 33, 616–623. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Mitroviç, Z.; Aurer, I.; Radman, I.; Ajdukoviç, R.; Sertiç, J.; Labar, B. FCγRIIIA and FCγRIIA Polymorphisms Are Not Associated with Response to Rituximab and CHOP in Patients with Diffuse Large B-Cell Lymphoma. Haematologica 2007, 92, 998–999. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Váróczy, L.; Zilahi, E.; Gyetvai, A.; Kajtár, B.; Gergely, L.; Sipka, S.; Illés, A. Fc-Gamma-Receptor IIIa Polymorphism and Gene Expression Profile Do Not Predict the Prognosis in Diffuse Large B-Cell Lymphoma Treated with R-CHOP Protocol. Pathol. Oncol. Res. 2012, 18, 43–48. [Google Scholar] [CrossRef]
  35. Danielou-Lazareth, A.; Henry, G.; Geromin, D.; Khaznadar, Z.; Briere, J.; Tamouza, R.; Cayuela, J.-M.; Thieblemont, C.; Toubert, A.; Dulphy, N. At Diagnosis, Diffuse Large B-Cell Lymphoma Patients Show Impaired Rituximab-Mediated NK-Cell Cytotoxicity. Eur. J. Immunol. 2013, 43, 1383–1388. [Google Scholar] [CrossRef] [Green Version]
  36. Cox, M.C.; Battella, S.; La Scaleia, R.; Pelliccia, S.; Di Napoli, A.; Porzia, A.; Cecere, F.; Alma, E.; Zingoni, A.; Mainiero, F.; et al. Tumor-Associated and Immunochemotherapy-Dependent Long-Term Alterations of the Peripheral Blood NK Cell Compartment in DLBCL Patients. Oncoimmunology 2015, 4, e990773. [Google Scholar] [CrossRef]
  37. Borgerding, A.; Hasenkamp, J.; Engelke, M.; Burkhart, N.; Trümper, L.; Wienands, J.; Glass, B. B-Lymphoma Cells Escape Rituximab-Triggered Elimination by NK Cells through Increased HLA Class I Expression. Exp. Hematol. 2010, 38, 213–221. [Google Scholar] [CrossRef]
  38. Dębska-Zielkowska, J.; Moszkowska, G.; Zieliński, M.; Zielińska, H.; Dukat-Mazurek, A.; Trzonkowski, P.; Stefańska, K. KIR Receptors as Key Regulators of NK Cells Activity in Health and Disease. Cells 2021, 10, 1777. [Google Scholar] [CrossRef]
  39. Makanga, D.R.; Jullien, M.; David, G.; Legrand, N.; Willem, C.; Dubreuil, L.; Walencik, A.; Touzeau, C.; Gastinne, T.; Tessoulin, B.; et al. Low Number of KIR Ligands in Lymphoma Patients Favors a Good Rituximab-Dependent NK Cell Response. Oncoimmunology 2021, 10, 1936392. [Google Scholar] [CrossRef]
  40. Beum, P.V.; Lindorfer, M.A.; Taylor, R.P. Within Peripheral Blood Mononuclear Cells, Antibody-Dependent Cellular Cytotoxicity of Rituximab-Opsonized Daudi Cells Is Promoted by NK Cells and Inhibited by Monocytes Due to Shaving. J. Immunol. 2008, 181, 2916–2924. [Google Scholar] [CrossRef] [Green Version]
  41. Beum, P.V.; Kennedy, A.D.; Williams, M.E.; Lindorfer, M.A.; Taylor, R.P. The Shaving Reaction: Rituximab/CD20 Complexes Are Removed from Mantle Cell Lymphoma and Chronic Lymphocytic Leukemia Cells by THP-1 Monocytes. J. Immunol. 2006, 176, 2600–2609. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Pedersen, A.E.; Jungersen, M.B.; Pedersen, C.D. Monocytes Mediate Shaving of B-Cell-Bound Anti-CD20 Antibodies. Immunology 2011, 133, 239–245. [Google Scholar] [CrossRef]
  43. Hansson, M.; Asea, A.; Ersson, U.; Hermodsson, S.; Hellstrand, K. Induction of Apoptosis in NK Cells by Monocyte-Derived Reactive Oxygen Metabolites. J. Immunol. 1996, 156, 42–47. Available online: https://www.jimmunol.org/content/156/1/42 (accessed on 30 October 2021). [PubMed]
  44. Klopotowska, M.; Bajor, M.; Graczyk-Jarzynka, A.; Kraft, A.; Pilch, Z.; Zhylko, A.; Firczuk, M.; Baranowska, I.; Lazniewski, M.; Plewczynski, D.; et al. PRDX-1 Supports the Survival and Antitumor Activity of Primary and CAR-Modified NK Cells under Oxidative Stress. Cancer Immunol. Res. 2021. [Google Scholar] [CrossRef] [PubMed]
  45. Werlenius, O.; Riise, R.E.; Simpanen, M.; Aurelius, J.; Thorén, F.B. CD20 Antibodies Induce Production and Release of Reactive Oxygen Species by Neutrophils. Blood 2014, 123, 4001–4002. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Honeychurch, J.; Alduaij, W.; Azizyan, M.; Cheadle, E.J.; Pelicano, H.; Ivanov, A.; Huang, P.; Cragg, M.S.; Illidge, T.M. Antibody-Induced Nonapoptotic Cell Death in Human Lymphoma and Leukemia Cells Is Mediated through a Novel Reactive Oxygen Species-Dependent Pathway. Blood 2012, 119, 3523–3533. [Google Scholar] [CrossRef]
  47. Werlenius, O.; Aurelius, J.; Hallner, A.; Akhiani, A.A.; Simpanen, M.; Martner, A.; Andersson, P.-O.; Hellstrand, K.; Thorén, F.B. Reactive Oxygen Species Induced by Therapeutic CD20 Antibodies Inhibit Natural Killer Cell-Mediated Antibody-Dependent Cellular Cytotoxicity against Primary CLL Cells. Oncotarget 2016, 7, 32046–32053. [Google Scholar] [CrossRef] [Green Version]
  48. de Charette, M.; Houot, R. Hide or Defend, the Two Strategies of Lymphoma Immune Evasion: Potential Implications for Immunotherapy. Haematologica 2018, 103, 1256–1268. [Google Scholar] [CrossRef]
  49. Upadhyay, R.; Hammerich, L.; Peng, P.; Brown, B.; Merad, M.; Brody, J.D. Lymphoma: Immune Evasion Strategies. Cancers 2015, 7, 736–762. [Google Scholar] [CrossRef] [Green Version]
  50. Rimsza, L.M.; Roberts, R.A.; Miller, T.P.; Unger, J.M.; LeBlanc, M.; Braziel, R.M.; Weisenberger, D.D.; Chan, W.C.; Muller-Hermelink, H.K.; Jaffe, E.S.; et al. Loss of MHC Class II Gene and Protein Expression in Diffuse Large B-Cell Lymphoma Is Related to Decreased Tumor Immunosurveillance and Poor Patient Survival Regardless of Other Prognostic Factors: A Follow-up Study from the Leukemia and Lymphoma Molecular Profiling Project. Blood 2004, 103, 4251–4258. [Google Scholar] [CrossRef] [Green Version]
  51. Challa-Malladi, M.; Lieu, Y.K.; Califano, O.; Holmes, A.; Bhagat, G.; Murty, V.V.; Dominguez-Sola, D.; Pasqualucci, L.; Dalla-Favera, R. Combined Genetic Inactivation of Beta2-Microglobulin and CD58 Reveals Frequent Escape from Immune Recognition in Diffuse Large B-Cell Lymphoma. Cancer Cell 2011, 20, 728–740. [Google Scholar] [CrossRef] [Green Version]
  52. Sordo-Bahamonde, C.; Lorenzo-Herrero, S.; Payer, Á.R.; Gonzalez, S.; López-Soto, A. Mechanisms of Apoptosis Resistance to NK Cell-Mediated Cytotoxicity in Cancer. Int. J. Mol. Sci. 2020, 21, 3726. [Google Scholar] [CrossRef]
  53. Matulis, S.M.; Boise, L.H. BCL2 Dependency in Diffuse Large B-Cell Lymphoma: It’s a Family Affair. Haematologica 2020, 105, 1993–1996. [Google Scholar] [CrossRef] [PubMed]
  54. Mestre-Escorihuela, C.; Rubio-Moscardo, F.; Richter, J.A.; Siebert, R.; Climent, J.; Fresquet, V.; Beltran, E.; Agirre, X.; Marugan, I.; Marín, M.; et al. Homozygous Deletions Localize Novel Tumor Suppressor Genes in B-Cell Lymphomas. Blood 2007, 109, 271–280. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Medema, J.P.; de Jong, J.; Peltenburg, L.T.; Verdegaal, E.M.; Gorter, A.; Bres, S.A.; Franken, K.L.; Hahne, M.; Albar, J.P.; Melief, C.J.; et al. Blockade of the Granzyme B/Perforin Pathway through Overexpression of the Serine Protease Inhibitor PI-9/SPI-6 Constitutes a Mechanism for Immune Escape by Tumors. Proc. Natl. Acad. Sci. USA 2001, 98, 11515–11520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Czuczman, M.S.; Olejniczak, S.; Gowda, A.; Kotowski, A.; Binder, A.; Kaur, H.; Knight, J.; Starostik, P.; Deans, J.; Hernandez-Ilizaliturri, F.J. Acquirement of Rituximab Resistance in Lymphoma Cell Lines Is Associated with Both Global CD20 Gene and Protein Down-Regulation Regulated at the Pretranscriptional and Posttranscriptional Levels. Clin. Cancer Res. 2008, 14, 1561–1570. [Google Scholar] [CrossRef] [Green Version]
  57. Jazirehi, A.R.; Vega, M.I.; Bonavida, B. Development of Rituximab-Resistant Lymphoma Clones with Altered Cell Signaling and Cross-Resistance to Chemotherapy. Cancer Res. 2007, 67, 1270–1281. [Google Scholar] [CrossRef] [Green Version]
  58. Li, B.; Zhao, L.; Guo, H.; Wang, C.; Zhang, X.; Wu, L.; Chen, L.; Tong, Q.; Qian, W.; Wang, H.; et al. Characterization of a Rituximab Variant with Potent Antitumor Activity against Rituximab-Resistant B-Cell Lymphoma. Blood 2009, 114, 5007–5015. [Google Scholar] [CrossRef] [Green Version]
  59. Aldeghaither, D.S.; Zahavi, D.J.; Murray, J.C.; Fertig, E.J.; Graham, G.T.; Zhang, Y.-W.; O’Connell, A.; Ma, J.; Jablonski, S.A.; Weiner, L.M. A Mechanism of Resistance to Antibody-Targeted Immune Attack. Cancer Immunol. Res. 2019, 7, 230–243. [Google Scholar] [CrossRef] [Green Version]
  60. Dubrovsky, L.; Pankov, D.; Brea, E.J.; Dao, T.; Scott, A.; Yan, S.; O’Reilly, R.J.; Liu, C.; Scheinberg, D.A. A TCR-Mimic Antibody to WT1 Bypasses Tyrosine Kinase Inhibitor Resistance in Human BCR-ABL+ Leukemias. Blood 2014, 123, 3296–3304. [Google Scholar] [CrossRef] [Green Version]
  61. Dubrovsky, L.; Brea, E.J.; Pankov, D.; Casey, E.; Dao, T.; Liu, C.; Scheinberg, D.A. Mechanisms of Leukemia Resistance to Antibody Dependent Cellular Cytotoxicity. Oncoimmunology 2016, 5, e1211221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Merkt, W.; Lorenz, H.-M.; Watzl, C. Rituximab Induces Phenotypical and Functional Changes of NK Cells in a Non-Malignant Experimental Setting. Arthritis Res. Ther. 2016, 18, 206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Capuano, C.; Romanelli, M.; Pighi, C.; Cimino, G.; Rago, A.; Molfetta, R.; Paolini, R.; Santoni, A.; Galandrini, R. Anti-CD20 Therapy Acts via FcγRIIIA to Diminish Responsiveness of Human Natural Killer Cells. Cancer Res. 2015, 75, 4097–4108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Wirt, T.; Rosskopf, S.; Rösner, T.; Eichholz, K.M.; Kahrs, A.; Lutz, S.; Kretschmer, A.; Valerius, T.; Klausz, K.; Otte, A.; et al. An Fc Double-Engineered CD20 Antibody with Enhanced Ability to Trigger Complement-Dependent Cytotoxicity and Antibody-Dependent Cell-Mediated Cytotoxicity. Transfus. Med. Hemother. 2017, 44, 292–300. [Google Scholar] [CrossRef]
  65. O’Nions, J.; Townsend, W. The Role of Obinutuzumab in the Management of Follicular Lymphoma. Future Oncol. 2019, 15, 3565–3578. [Google Scholar] [CrossRef]
  66. Terszowski, G.; Klein, C.; Stern, M. KIR/HLA Interactions Negatively Affect Rituximab- but Not GA101 (Obinutuzumab)-Induced Antibody-Dependent Cellular Cytotoxicity. J. Immunol. 2014, 192, 5618–5624. [Google Scholar] [CrossRef] [Green Version]
  67. Binyamin, L.; Alpaugh, R.K.; Hughes, T.L.; Lutz, C.T.; Campbell, K.S.; Weiner, L.M. Blocking NK Cell Inhibitory Self-Recognition Promotes Antibody-Dependent Cellular Cytotoxicity in a Model of Anti-Lymphoma Therapy. J. Immunol. 2008, 180, 6392–6401. [Google Scholar] [CrossRef]
  68. Sehn, L.H.; Martelli, M.; Trněný, M.; Liu, W.; Bolen, C.R.; Knapp, A.; Sahin, D.; Sellam, G.; Vitolo, U. A Randomized, Open-Label, Phase III Study of Obinutuzumab or Rituximab plus CHOP in Patients with Previously Untreated Diffuse Large B-Cell Lymphoma: Final Analysis of GOYA. J. Hematol. Oncol. 2020, 13, 71. [Google Scholar] [CrossRef]
  69. Le Gouill, S.; Ghesquières, H.; Oberic, L.; Morschhauser, F.; Tilly, H.; Ribrag, V.; Lamy, T.; Thieblemont, C.; Maisonneuve, H.; Gressin, R.; et al. Obinutuzumab vs Rituximab for Advanced DLBCL: A PET-Guided and Randomized Phase 3 Study by LYSA. Blood 2021, 137, 2307–2320. [Google Scholar] [CrossRef]
  70. Marcus, R.; Davies, A.; Ando, K.; Klapper, W.; Opat, S.; Owen, C.; Phillips, E.; Sangha, R.; Schlag, R.; Seymour, J.F.; et al. Obinutuzumab for the First-Line Treatment of Follicular Lymphoma. N. Engl. J. Med. 2017, 377, 1331–1344. [Google Scholar] [CrossRef]
  71. Goede, V.; Fischer, K.; Busch, R.; Engelke, A.; Eichhorst, B.; Wendtner, C.M.; Chagorova, T.; de la Serna, J.; Dilhuydy, M.-S.; Illmer, T.; et al. Obinutuzumab plus Chlorambucil in Patients with CLL and Coexisting Conditions. N. Engl. J. Med. 2014, 370, 1101–1110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Wayne, J.L.; Campbell, M.-A.; Marulappa, S.; Jain, V. Abstract 3784: AME-133v, a Humanized, Fc-Engineered Anti-CD20 Monoclonal Antibody, Demonstrates Greater B Cell Depletion than Rituxan in Vitro. Cancer Res. 2012, 72, 3784. [Google Scholar] [CrossRef]
  73. Wayne, J.L.; Ganjoo, K.N.; Pohlman, B.L.; De Vos, S.; Flinn, I.W.; Dang, N.H.; Mapara, M.Y.; Smith, M.R.; O’Reilly, A.M.; Marulappa, S.Y.; et al. Efficacy of Ocaratuzumab (AME-133v) in Relapsed Follicular Lymphoma Patients Refractory to Prior Rituximab. J. Clin. Oncol. 2012, 30, 8081. [Google Scholar] [CrossRef]
  74. Forero-Torres, A.; de Vos, S.; Pohlman, B.L.; Pashkevich, M.; Cronier, D.M.; Dang, N.H.; Carpenter, S.P.; Allan, B.W.; Nelson, J.G.; Slapak, C.A.; et al. Results of a Phase 1 Study of AME-133v (LY2469298), an Fc-Engineered Humanized Monoclonal Anti-CD20 Antibody, in FcγRIIIa-Genotyped Patients with Previously Treated Follicular Lymphoma. Clin. Cancer Res. 2012, 18, 1395–1403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Ganjoo, K.N.; de Vos, S.; Pohlman, B.L.; Flinn, I.W.; Forero-Torres, A.; Enas, N.H.; Cronier, D.M.; Dang, N.H.; Foon, K.A.; Carpenter, S.P.; et al. Phase 1/2 Study of Ocaratuzumab, an Fc-Engineered Humanized Anti-CD20 Monoclonal Antibody, in Low-Affinity FcγRIIIa Patients with Previously Treated Follicular Lymphoma. Leuk. Lymphoma 2015, 56, 42–48. [Google Scholar] [CrossRef]
  76. Horton, H.M.; Bernett, M.J.; Pong, E.; Peipp, M.; Karki, S.; Chu, S.Y.; Richards, J.O.; Vostiar, I.; Joyce, P.F.; Repp, R.; et al. Potent In Vitro and In Vivo Activity of an Fc-Engineered Anti-CD19 Monoclonal Antibody against Lymphoma and Leukemia. Cancer Res. 2008, 68, 8049–8057. [Google Scholar] [CrossRef] [Green Version]
  77. Jurczak, W.; Zinzani, P.L.; Gaidano, G.; Goy, A.; Provencio, M.; Nagy, Z.; Robak, T.; Maddocks, K.; Buske, C.; Ambarkhane, S.; et al. Phase IIa Study of the CD19 Antibody MOR208 in Patients with Relapsed or Refractory B-Cell Non-Hodgkin’s Lymphoma. Ann. Oncol. 2018, 29, 1266–1272. [Google Scholar] [CrossRef]
  78. Rosario, M.; Bai, L.; Kong, L.; Collins, L.I.; Schneider, S.E.; Chen, X.; Han, K.; Jeng, E.K.; Rhode, P.R.; Leong, J.W.; et al. The IL-15-Based ALT-803 Complex Enhances FcγRIIIa-Triggered NK Cell Responses and in Vivo Clearance of B Cell Lymphomas. Clin. Cancer Res. 2016, 22, 596–608. [Google Scholar] [CrossRef] [Green Version]
  79. Bittenbring, J.T.; Neumann, F.; Altmann, B.; Achenbach, M.; Reichrath, J.; Ziepert, M.; Geisel, J.; Regitz, E.; Held, G.; Pfreundschuh, M. Vitamin D Deficiency Impairs Rituximab-Mediated Cellular Cytotoxicity and Outcome of Patients with Diffuse Large B-Cell Lymphoma Treated with but Not without Rituximab. J. Clin. Oncol. 2014, 32, 3242–3248. [Google Scholar] [CrossRef]
  80. Neumann, F.; Acker, F.; Schormann, C.; Pfreundschuh, M.; Bittenbring, J.T. Determination of Optimum Vitamin D3 Levels for NK Cell-Mediated Rituximab- and Obinutuzumab-Dependent Cellular Cytotoxicity. Cancer Immunol. Immunother. 2018, 67, 1709–1718. [Google Scholar] [CrossRef]
  81. Pierpont, T.M.; Limper, C.B.; Richards, K.L. Past, Present, and Future of Rituximab—The World’s First Oncology Monoclonal Antibody Therapy. Front. Oncol. 2018, 8, 163. [Google Scholar] [CrossRef] [PubMed]
  82. Felberg, A.; Taszner, M.; Urban, A.; Majeranowski, A.; Jaskuła, K.; Jurkiewicz, A.; Stasiłojć, G.; Blom, A.M.; Zaucha, J.M.; Okrój, M. Monitoring of the Complement System Status in Patients With B-Cell Malignancies Treated with Rituximab. Front. Immunol. 2020, 11, 2956. [Google Scholar] [CrossRef] [PubMed]
  83. Bordron, A.; Bagacean, C.; Mohr, A.; Tempescul, A.; Bendaoud, B.; Deshayes, S.; Dalbies, F.; Buors, C.; Saad, H.; Berthou, C.; et al. Resistance to Complement Activation, Cell Membrane Hypersialylation and Relapses in Chronic Lymphocytic Leukemia Patients Treated with Rituximab and Chemotherapy. Oncotarget 2018, 9, 31590–31605. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Kennedy, A.D.; Beum, P.V.; Solga, M.D.; DiLillo, D.J.; Lindorfer, M.A.; Hess, C.E.; Densmore, J.J.; Williams, M.E.; Taylor, R.P. Rituximab Infusion Promotes Rapid Complement Depletion and Acute CD20 Loss in Chronic Lymphocytic Leukemia. J. Immunol. 2004, 172, 3280–3288. [Google Scholar] [CrossRef] [Green Version]
  85. Füst, G.; Czink, E.; Minh, D.; Miszlay, Z.; Varga, L.; Hollán, S.R. Depressed Classical Complement Pathway Activities in Chronic Lymphocytic Leukaemia. Clin. Exp. Immunol. 1985, 60, 489–495. [Google Scholar] [PubMed]
  86. Füst, G.; Miszlay, Z.s.; Czink, E.; Varga, L.; Pálóczi, K.; Szegedi, G.; Hollán, S.R. C1 and C4 Abnormalities in Chronic Lymphocytic Leukaemia and Their Significance. Immunol. Lett. 1987, 14, 255–259. [Google Scholar] [CrossRef]
  87. Heath, M.E.; Cheson, B.D. Defective Complement Activity in Chronic Lymphocytic Leukemia. Am. J. Hematol. 1985, 19, 63–73. [Google Scholar] [CrossRef]
  88. Middleton, O.; Cosimo, E.; Dobbin, E.; McCaig, A.M.; Clarke, C.; Brant, A.M.; Leach, M.T.; Michie, A.M.; Wheadon, H. Complement Deficiencies Limit CD20 Monoclonal Antibody Treatment Efficacy in CLL. Leukemia 2015, 29, 107–114. [Google Scholar] [CrossRef] [Green Version]
  89. Klepfish, A.; Rachmilewitz, E.A.; Kotsianidis, I.; Patchenko, P.; Schattner, A. Adding Fresh Frozen Plasma to Rituximab for the Treatment of Patients with Refractory Advanced CLL. QJM Int. J. Med. 2008, 101, 737–740. [Google Scholar] [CrossRef] [Green Version]
  90. Golay, J.; Lazzari, M.; Facchinetti, V.; Bernasconi, S.; Borleri, G.; Barbui, T.; Rambaldi, A.; Introna, M. CD20 Levels Determine the in Vitro Susceptibility to Rituximab and Complement of B-Cell Chronic Lymphocytic Leukemia: Further Regulation by CD55 and CD59. Blood 2001, 98, 3383–3389. [Google Scholar] [CrossRef] [Green Version]
  91. Winkler, M.T.; Bushey, R.T.; Gottlin, E.B.; Campa, M.J.; Guadalupe, E.S.; Volkheimer, A.D.; Weinberg, J.B.; Patz, E.F. Enhanced CDC of B Cell Chronic Lymphocytic Leukemia Cells Mediated by Rituximab Combined with a Novel Anti-Complement Factor H Antibody. PLoS ONE 2017, 12, e0179841. [Google Scholar] [CrossRef] [Green Version]
  92. Gaetano, N.D.; Cittera, E.; Nota, R.; Vecchi, A.; Grieco, V.; Scanziani, E.; Botto, M.; Introna, M.; Golay, J. Complement Activation Determines the Therapeutic Activity of Rituximab In Vivo. J. Immunol. 2003, 171, 1581–1587. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Gong, Q.; Ou, Q.; Ye, S.; Lee, W.P.; Cornelius, J.; Diehl, L.; Lin, W.Y.; Hu, Z.; Lu, Y.; Chen, Y.; et al. Importance of Cellular Microenvironment and Circulatory Dynamics in B Cell Immunotherapy. J. Immunol. 2005, 174, 817–826. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Song, G.; Song, G.; Ni, H.; Gu, L.; Liu, H.; Chen, B.; He, B.; Pan, Y.; Wang, S.; Cho, W.C. Deregulated Expression of MiR-224 and Its Target Gene: CD59 Predicts Outcome of Diffuse Large B-Cell Lymphoma Patients Treated with R-CHOP. Curr. Cancer Drug Targets 2014, 14, 659–670. [Google Scholar] [CrossRef] [PubMed]
  95. Sugita, Y.; Masuho, Y. CD59: Its Role in Complement Regulation and Potential for Therapeutic Use. Immunotechnology 1995, 1, 157–168. [Google Scholar] [CrossRef]
  96. Wang, S.-Y.; Veeramani, S.; Racila, E.; Cagley, J.; Fritzinger, D.C.; Vogel, C.-W.; St John, W.; Weiner, G.J. Depletion of the C3 Component of Complement Enhances the Ability of Rituximab-Coated Target Cells to Activate Human NK Cells and Improves the Efficacy of Monoclonal Antibody Therapy in an in Vivo Model. Blood 2009, 114, 5322–5330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Wang, S.-Y.; Racila, E.; Taylor, R.P.; Weiner, G.J. NK-Cell Activation and Antibody-Dependent Cellular Cytotoxicity Induced by Rituximab-Coated Target Cells Is Inhibited by the C3b Component of Complement. Blood 2008, 111, 1456–1463. [Google Scholar] [CrossRef]
  98. Racila, E.; Link, B.K.; Weng, W.-K.; Witzig, T.E.; Ansell, S.; Maurer, M.J.; Huang, J.; Dahle, C.; Halwani, A.; Levy, R.; et al. A Polymorphism in the Complement Component C1qA Correlates with Prolonged Response Following Rituximab Therapy of Follicular Lymphoma. Clin. Cancer Res. 2008, 14, 6697–6703. [Google Scholar] [CrossRef] [Green Version]
  99. Jin, X.; Ding, H.; Ding, N.; Fu, Z.; Song, Y.; Zhu, J. Homozygous A Polymorphism of the Complement C1qA276correlates with Prolonged Overall Survival in Patients with Diffuse Large B Cell Lymphoma Treated with R-CHOP. J. Hematol. Oncol. 2012, 5, 51. [Google Scholar] [CrossRef] [Green Version]
  100. Deans, J.P.; Robbins, S.M.; Polyak, M.J.; Savage, J.A. Rapid Redistribution of CD20 to a Low Density Detergent-Insoluble Membrane Compartment. J. Biol. Chem. 1998, 273, 344–348. [Google Scholar] [CrossRef] [Green Version]
  101. Rezvani, A.R.; Maloney, D.G. Rituximab Resistance. Best Pract. Res. Clin. Haematol. 2011, 24, 203–216. [Google Scholar] [CrossRef] [PubMed]
  102. Mollinedo, F.; Gajate, C. Lipid Rafts as Signaling Hubs in Cancer Cell Survival/Death and Invasion: Implications in Tumor Progression and Therapy: Thematic Review Series: Biology of Lipid Rafts. J. Lipid Res. 2020, 61, 611–635. [Google Scholar] [CrossRef] [Green Version]
  103. Hammadi, M.; Youinou, P.; Tempescul, A.; Tobón, G.; Berthou, C.; Bordron, A.; Pers, J.-O. Membrane Microdomain Sphingolipids Are Required for Anti-CD20-Induced Death of Chronic Lymphocytic Leukemia B Cells. Haematologica 2012, 97, 288–296. [Google Scholar] [CrossRef] [Green Version]
  104. Meyer zum Büschenfelde, C.; Feuerstacke, Y.; Götze, K.S.; Scholze, K.; Peschel, C. GM1 Expression of Non-Hodgkin’s Lymphoma Determines Susceptibility to Rituximab Treatment. Cancer Res. 2008, 68, 5414–5422. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Meri, S. Self-Nonself Discrimination by the Complement System. FEBS Lett. 2016, 590, 2418–2434. [Google Scholar] [CrossRef] [Green Version]
  106. Boyd, R.S.; Jukes-Jones, R.; Walewska, R.; Brown, D.; Dyer, M.J.S.; Cain, K. Protein Profiling of Plasma Membranes Defines Aberrant Signaling Pathways in Mantle Cell Lymphoma. Mol. Cell Proteom. 2009, 8, 1501–1515. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Hirpara, J.L.; Loh, T.; Ng, S.B.; Chng, W.J.; Pervaiz, S. Aberrant Localization of Apoptosis Protease Activating Factor-1 in Lipid Raft Sub-Domains of Diffuse Large B Cell Lymphomas. Oncotarget 2016, 7, 83964–83975. [Google Scholar] [CrossRef] [Green Version]
  108. Winiarska, M.; Bil, J.; Wilczek, E.; Wilczynski, G.M.; Lekka, M.; Engelberts, P.J.; Mackus, W.J.M.; Gorska, E.; Bojarski, L.; Stoklosa, T.; et al. Statins Impair Antitumor Effects of Rituximab by Inducing Conformational Changes of CD20. PLoS Med. 2008, 5, e64. [Google Scholar] [CrossRef] [Green Version]
  109. Gisselbrecht, C.; Glass, B.; Mounier, N.; Singh Gill, D.; Linch, D.C.; Trneny, M.; Bosly, A.; Ketterer, N.; Shpilberg, O.; Hagberg, H.; et al. Salvage Regimens With Autologous Transplantation for Relapsed Large B-Cell Lymphoma in the Rituximab Era. J. Clin. Oncol. 2010, 28, 4184–4190. [Google Scholar] [CrossRef] [Green Version]
  110. Martín, A.; Conde, E.; Arnan, M.; Canales, M.A.; Deben, G.; Sancho, J.M.; Andreu, R.; Salar, A.; García-Sanchez, P.; Vázquez, L.; et al. R-ESHAP as Salvage Therapy for Patients with Relapsed or Refractory Diffuse Large B-Cell Lymphoma: The Influence of Prior Exposure to Rituximab on Outcome. A GEL/TAMO Study. Haematologica 2008, 93, 1829–1836. [Google Scholar] [CrossRef] [Green Version]
  111. Barth, M.J.; Hernandez-Ilizaliturri, F.J.; Mavis, C.; Tsai, P.-C.; Gibbs, J.F.; Deeb, G.; Czuczman, M.S. Ofatumumab Demonstrates Activity against Rituximab-Sensitive and -Resistant Cell Lines, Lymphoma Xenografts and Primary Tumour Cells from Patients with B-Cell Lymphoma. Br. J. Haematol. 2012, 156, 490–498. [Google Scholar] [CrossRef]
  112. Takei, K.; Yamazaki, T.; Sawada, U.; Ishizuka, H.; Aizawa, S. Analysis of Changes in CD20, CD55, and CD59 Expression on Established Rituximab-Resistant B-Lymphoma Cell Lines. Leuk. Res. 2006, 30, 625–631. [Google Scholar] [CrossRef]
  113. Tsai, P.-C.; Hernandez-Ilizaliturri, F.J.; Bangia, N.; Olejniczak, S.H.; Czuczman, M.S. Regulation of CD20 in Rituximab-Resistant Cell Lines and B-Cell Non-Hodgkin Lymphoma. Clin. Cancer Res. 2012, 18, 1039–1050. [Google Scholar] [CrossRef] [Green Version]
  114. Golay, J.; Zaffaroni, L.; Vaccari, T.; Lazzari, M.; Borleri, G.M.; Bernasconi, S.; Tedesco, F.; Rambaldi, A.; Introna, M. Biologic Response of B Lymphoma Cells to Anti-CD20 Monoclonal Antibody Rituximab in Vitro: CD55 and CD59 Regulate Complement-Mediated Cell Lysis. Blood 2000, 95, 3900–3908. [Google Scholar] [CrossRef]
  115. Hiraga, J.; Tomita, A.; Sugimoto, T.; Shimada, K.; Ito, M.; Nakamura, S.; Kiyoi, H.; Kinoshita, T.; Naoe, T. Down-Regulation of CD20 Expression in B-Cell Lymphoma Cells after Treatment with Rituximab-Containing Combination Chemotherapies: Its Prevalence and Clinical Significance. Blood 2009, 113, 4885–4893. [Google Scholar] [CrossRef] [Green Version]
  116. Davis, T.A.; Czerwinski, D.K.; Levy, R. Therapy of B-Cell Lymphoma with Anti-CD20 Antibodies Can Result in the Loss of CD20 Antigen Expression. Clin. Cancer Res. 1999, 5, 611–615. [Google Scholar]
  117. Haidar, J.H.; Shamseddine, A.; Salem, Z.; Mrad, Y.A.; Nasr, M.R.; Zaatari, G.; Bazarbachi, A. Loss of CD20 Expression in Relapsed Lymphomas after Rituximab Therapy. Eur. J. Haematol. 2003, 70, 330–332. [Google Scholar] [CrossRef]
  118. Rawal, Y.B.; Nuovo, G.J.; Frambach, G.E.; Porcu, P.; Baiocchi, R.A.; Magro, C.M. The Absence of CD20 Messenger RNA in Recurrent Cutaneous B-Cell Lymphoma Following Rituximab Therapy. J. Cutan. Pathol. 2005, 32, 616–621. [Google Scholar] [CrossRef]
  119. Duman, B.B.; Sahin, B.; Ergin, M.; Guvenc, B. Loss of CD20 Antigen Expression after Rituximab Therapy of CD20 Positive B Cell Lymphoma (Diffuse Large B Cell Extranodal Marginal Zone Lymphoma Combination): A Case Report and Review of the Literature. Med. Oncol. 2012, 29, 1223–1226. [Google Scholar] [CrossRef]
  120. van Meerten, T.; Hagenbeek, A. CD20-Targeted Therapy: A Breakthrough in the Treatment of Non-Hodgkin’s Lymphoma. Neth. J. Med. 2009, 67, 251–259. [Google Scholar]
  121. Bobrowicz, M.; Dwojak, M.; Pyrzynska, B.; Stachura, J.; Muchowicz, A.; Berthel, E.; Dalla-Venezia, N.; Kozikowski, M.; Siernicka, M.; Miazek, N.; et al. HDAC6 Inhibition Upregulates CD20 Levels and Increases the Efficacy of Anti-CD20 Monoclonal Antibodies. Blood 2017, 130, 1628–1638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Li, H.; Ayer, L.M.; Lytton, J.; Deans, J.P. Store-Operated Cation Entry Mediated by CD20 in Membrane Rafts. J. Biol. Chem. 2003, 278, 42427–42434. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Kozlova, V.; Ledererova, A.; Ladungova, A.; Peschelova, H.; Janovska, P.; Slusarczyk, A.; Domagala, J.; Kopcil, P.; Vakulova, V.; Oppelt, J.; et al. CD20 Is Dispensable for B-Cell Receptor Signaling but Is Required for Proper Actin Polymerization, Adhesion and Migration of Malignant B Cells. PLoS ONE 2020, 15, e0229170. [Google Scholar] [CrossRef] [PubMed]
  124. Pyrzynska, B.; Dwojak, M.; Zerrouqi, A.; Morlino, G.; Zapala, P.; Miazek, N.; Zagozdzon, A.; Bojarczuk, K.; Bobrowicz, M.; Siernicka, M.; et al. FOXO1 Promotes Resistance of Non-Hodgkin Lymphomas to Anti-CD20-Based Therapy. Oncoimmunology 2018, 7, e1423183. [Google Scholar] [CrossRef] [Green Version]
  125. Winiarska, M.; Bojarczuk, K.; Pyrzynska, B.; Bil, J.; Siernicka, M.; Dwojak, M.; Bobrowicz, M.; Miazek, N.; Zapala, P.; Zagozdzon, A.; et al. Inhibitors of SRC Kinases Impair Antitumor Activity of Anti-CD20 Monoclonal Antibodies. mAbs 2014, 6, 1300–1313. [Google Scholar] [CrossRef] [Green Version]
  126. Winiarska, M.; Nowis, D.; Bil, J.; Glodkowska-Mrowka, E.; Muchowicz, A.; Wanczyk, M.; Bojarczuk, K.; Dwojak, M.; Firczuk, M.; Wilczek, E.; et al. Prenyltransferases Regulate CD20 Protein Levels and Influence Anti-CD20 Monoclonal Antibody-Mediated Activation of Complement-Dependent Cytotoxicity. J. Biol. Chem. 2012, 287, 31983–31993. [Google Scholar] [CrossRef] [Green Version]
  127. Bil, J.; Winiarska, M.; Nowis, D.; Bojarczuk, K.; Dabrowska-Iwanicka, A.; Basak, G.W.; Sułek, K.; Jakobisiak, M.; Golab, J. Bortezomib Modulates Surface CD20 in B-Cell Malignancies and Affects Rituximab-Mediated Complement-Dependent Cytotoxicity. Blood 2010, 115, 3745–3755. [Google Scholar] [CrossRef] [Green Version]
  128. Winiarska, M.; Bil, J.; Nowis, D.; Golab, J. Proteolytic Pathways Involved in Modulation of CD20 Levels. Autophagy 2010, 6, 810–812. [Google Scholar] [CrossRef] [Green Version]
  129. Manshouri, T.; Do, K.; Wang, X.; Giles, F.J.; O’Brien, S.M.; Saffer, H.; Thomas, D.; Jilani, I.; Kantarjian, H.M.; Keating, M.J.; et al. Circulating CD20 Is Detectable in the Plasma of Patients with Chronic Lymphocytic Leukemia and Is of Prognostic Significance. Blood 2003, 101, 2507–2513. [Google Scholar] [CrossRef]
  130. Aung, T.; Chapuy, B.; Vogel, D.; Wenzel, D.; Oppermann, M.; Lahmann, M.; Weinhage, T.; Menck, K.; Hupfeld, T.; Koch, R.; et al. Exosomal Evasion of Humoral Immunotherapy in Aggressive B-Cell Lymphoma Modulated by ATP-Binding Cassette Transporter A3. Proc. Natl. Acad. Sci. USA 2011, 108, 15336–15341. [Google Scholar] [CrossRef] [Green Version]
  131. Marshall, M.J.E.; Stopforth, R.J.; Cragg, M.S. Therapeutic Antibodies: What Have We Learnt from Targeting CD20 and Where Are We Going? Front. Immunol. 2017, 8, 1245. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. de Jong, M.R.W.; Visser, L.; Huls, G.; Diepstra, A.; van Vugt, M.; Ammatuna, E.; van Rijn, R.S.; Vellenga, E.; van den Berg, A.; Fehrmann, R.S.N.; et al. Identification of Relevant Drugable Targets in Diffuse Large B-Cell Lymphoma Using a Genome-Wide Unbiased CD20 Guilt-by Association Approach. PLoS ONE 2018, 13, e0193098. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Diebolder, C.A.; Beurskens, F.J.; de Jong, R.N.; Koning, R.I.; Strumane, K.; Lindorfer, M.A.; Voorhorst, M.; Ugurlar, D.; Rosati, S.; Heck, A.J.R.; et al. Complement Is Activated by IgG Hexamers Assembled at the Cell Surface. Science 2014, 343, 1260–1263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. de Jong, R.N.; Beurskens, F.J.; Verploegen, S.; Strumane, K.; van Kampen, M.D.; Voorhorst, M.; Horstman, W.; Engelberts, P.J.; Oostindie, S.C.; Wang, G.; et al. A Novel Platform for the Potentiation of Therapeutic Antibodies Based on Antigen-Dependent Formation of IgG Hexamers at the Cell Surface. PLoS Biol. 2016, 14, e1002344. [Google Scholar] [CrossRef] [Green Version]
  135. Olejniczak, S.H.; Hernandez-Ilizaliturri, F.J.; Clements, J.L.; Czuczman, M.S. Acquired Resistance to Rituximab Is Associated with Chemotherapy Resistance Resulting from Decreased Bax and Bak Expression. Clin. Cancer Res. 2008, 14, 1550–1560. [Google Scholar] [CrossRef] [Green Version]
  136. Bonavida, B. Rituximab-Induced Inhibition of Antiapoptotic Cell Survival Pathways: Implications in Chemo/Immunoresistance, Rituximab Unresponsiveness, Prognostic and Novel Therapeutic Interventions. Oncogene 2007, 26, 3629–3636. [Google Scholar] [CrossRef] [Green Version]
  137. Liu, Y.; Gu, J.J.; Yang, L.; Tsai, P.-C.; Guo, Y.; Xue, K.; Xia, Z.; Liu, X.; Lv, F.; Cao, J.; et al. The Adhesion Molecule ICAM-1 in Diffuse Large B-Cell Lymphoma Post-Rituximab Era: Relationship with Prognostic Importance and Rituximab Resistance. Aging (Albany NY) 2020, 13, 181–193. [Google Scholar] [CrossRef]
  138. Gu, J.J.; Singh, A.; Xue, K.; Mavis, C.; Barth, M.; Yanamadala, V.; Lenz, P.; Grau, M.; Lenz, G.; Czuczman, M.S.; et al. Up-Regulation of Hexokinase II Contributes to Rituximab-Chemotherapy Resistance and Is a Clinically Relevant Target for Therapeutic Development. Oncotarget 2018, 9, 4020–4033. [Google Scholar] [CrossRef] [Green Version]
  139. Hanahan, D.; Weinberg, R.A. Hallmarks of Cancer: The Next Generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [Green Version]
  140. .Sircar, A.; Chowdhury, S.M.; Hart, A.; Bell, W.C.; Singh, S.; Sehgal, L.; Epperla, N. Impact and Intricacies of Bone Marrow Microenvironment in B-Cell Lymphomas: From Biology to Therapy. Int. J. Mol. Sci. 2020, 21, 904. [Google Scholar] [CrossRef] [Green Version]
  141. Marquez, M.-E.; Hernández-Uzcátegui, O.; Cornejo, A.; Vargas, P.; Da Costa, O. Bone Marrow Stromal Mesenchymal Cells Induce down Regulation of CD20 Expression on B-CLL: Implications for Rituximab Resistance in CLL. Br. J. Haematol. 2015, 169, 211–218. [Google Scholar] [CrossRef] [PubMed]
  142. Mraz, M.; Zent, C.S.; Church, A.K.; Jelinek, D.F.; Wu, X.; Pospisilova, S.; Ansell, S.M.; Novak, A.J.; Kay, N.E.; Witzig, T.E.; et al. Bone Marrow Stromal Cells Protect Lymphoma B-Cells from Rituximab-Induced Apoptosis and Targeting Integrin α-4-β-1 (VLA-4) with Natalizumab Can Overcome This Resistance. Br. J. Haematol. 2011, 155, 53–64. [Google Scholar] [CrossRef] [PubMed]
  143. Zhong, W.; Zhu, Z.; Xu, X.; Zhang, H.; Xiong, H.; Li, Q.; Wei, Y. Human Bone Marrow-Derived Mesenchymal Stem Cells Promote the Growth and Drug-Resistance of Diffuse Large B-Cell Lymphoma by Secreting IL-6 and Elevating IL-17A Levels. J. Exp. Clin. Cancer Res. 2019, 38, 73. [Google Scholar] [CrossRef] [PubMed]
  144. Zhong, W.; Li, Q. Rituximab or Irradiation Promotes IL-17 Secretion and Thereby Induces Resistance to Rituximab or Irradiation. Cell. Mol. Immunol. 2017, 14, 1020–1022. [Google Scholar] [CrossRef] [PubMed]
  145. Abès, R.; Gélizé, E.; Fridman, W.H.; Teillaud, J.-L. Long-Lasting Antitumor Protection by Anti-CD20 Antibody through Cellular Immune Response. Blood 2010, 116, 926–934. [Google Scholar] [CrossRef] [Green Version]
  146. Deligne, C.; Metidji, A.; Fridman, W.-H.; Teillaud, J.-L. Anti-CD20 Therapy Induces a Memory Th1 Response through the IFN-γ/IL-12 Axis and Prevents Protumor Regulatory T-Cell Expansion in Mice. Leukemia 2015, 29, 947–957. [Google Scholar] [CrossRef]
  147. Opinto, G.; Vegliante, M.C.; Negri, A.; Skrypets, T.; Loseto, G.; Pileri, S.A.; Guarini, A.; Ciavarella, S. The Tumor Microenvironment of DLBCL in the Computational Era. Front. Oncol. 2020, 10, 351. [Google Scholar] [CrossRef] [Green Version]
  148. Lenz, G.; Wright, G.; Dave, S.S.; Xiao, W.; Powell, J.; Zhao, H.; Xu, W.; Tan, B.; Goldschmidt, N.; Iqbal, J.; et al. Stromal Gene Signatures in Large-B-Cell Lymphomas. N. Engl. J. Med. 2008, 359, 2313–2323. [Google Scholar] [CrossRef] [Green Version]
  149. Marchesi, F.; Cirillo, M.; Bianchi, A.; Gately, M.; Olimpieri, O.M.; Cerchiara, E.; Renzi, D.; Micera, A.; Balzamino, B.O.; Bonini, S.; et al. High Density of CD68+/CD163+ Tumour-Associated Macrophages (M2-TAM) at Diagnosis Is Significantly Correlated to Unfavorable Prognostic Factors and to Poor Clinical Outcomes in Patients with Diffuse Large B-Cell Lymphoma. Hematol. Oncol. 2015, 33, 110–112. [Google Scholar] [CrossRef]
  150. Desai, M.; Newberry, K.; Ou, Z.; Wang, M.; Zhang, L. Lenalidomide in Relapsed or Refractory Mantle Cell Lymphoma: Overview and Perspective. Ther. Adv. Hematol. 2014, 5, 91–101. [Google Scholar] [CrossRef]
  151. Witzig, T.E.; Vose, J.M.; Zinzani, P.L.; Reeder, C.B.; Buckstein, R.; Polikoff, J.A.; Bouabdallah, R.; Haioun, C.; Tilly, H.; Guo, P.; et al. An International Phase II Trial of Single-Agent Lenalidomide for Relapsed or Refractory Aggressive B-Cell Non-Hodgkin’s Lymphoma. Ann. Oncol. 2011, 22, 1622–1627. [Google Scholar] [CrossRef] [PubMed]
  152. Fowler, N.H.; Davis, R.E.; Rawal, S.; Nastoupil, L.; Hagemeister, F.B.; McLaughlin, P.; Kwak, L.W.; Romaguera, J.E.; Fanale, M.A.; Fayad, L.E.; et al. Safety, Activity, and Immune Effects of Lenalidomide and Rituximab in Untreated Indolent Lymphoma. Lancet Oncol. 2014, 15, 1311–1318. [Google Scholar] [CrossRef] [Green Version]
  153. Gribben, J.G.; Fowler, N.; Morschhauser, F. Mechanisms of Action of Lenalidomide in B-Cell Non-Hodgkin Lymphoma. J. Clin. Oncol. 2015, 33, 2803–2811. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Walewski, J. Novel Monoclonal Antibodies for Diffuse Large B-Cell Lymphoma. Acta Haematol. Pol. 2021, 52, 329–333. [Google Scholar] [CrossRef]
  155. Bhat, S.A.; Czuczman, M.S. Novel Antibodies in the Treatment of Non-Hodgkin’s Lymphoma. Neth. J. Med. 2009, 67, 311–321. [Google Scholar]
  156. Patriarca, A.; Gaidano, G. Investigational Drugs for the Treatment of Diffuse Large B-Cell Lymphoma. Expert Opin. Investig. Drugs 2021, 30, 25–38. [Google Scholar] [CrossRef]
  157. Bobrowicz, M.; Kubacz, M.; Slusarczyk, A.; Winiarska, M. CD37 in B Cell Derived Tumors—More than Just a Docking Point for Monoclonal Antibodies. Int. J. Mol. Sci. 2020, 21, 9531. [Google Scholar] [CrossRef]
  158. Capuano, C.; Pighi, C.; Battella, S.; De Federicis, D.; Galandrini, R.; Palmieri, G. Harnessing CD16-Mediated NK Cell Functions to Enhance Therapeutic Efficacy of Tumor-Targeting MAbs. Cancers 2021, 13, 2500. [Google Scholar] [CrossRef]
  159. Viardot, A.; Goebeler, M.-E.; Hess, G.; Neumann, S.; Pfreundschuh, M.; Adrian, N.; Zettl, F.; Libicher, M.; Sayehli, C.; Stieglmaier, J.; et al. Phase 2 Study of the Bispecific T-Cell Engager (BiTE) Antibody Blinatumomab in Relapsed/Refractory Diffuse Large B-Cell Lymphoma. Blood 2016, 127, 1410–1416. [Google Scholar] [CrossRef]
  160. Viardot, A.; Hess, G.; Bargou, R.C.; Morley, N.J.; Gritti, G.; Goebeler, M.-E.; Iskander, K.; Cohan, D.; Zhang, A.; Franklin, J.; et al. Durability of Complete Response after Blinatumomab Therapy for Relapsed/Refractory Diffuse Large B-Cell Lymphoma. Leuk. Lymphoma 2020, 61, 2767–2770. [Google Scholar] [CrossRef]
  161. Dufner, V.; Sayehli, C.M.; Chatterjee, M.; Hummel, H.D.; Gelbrich, G.; Bargou, R.C.; Goebeler, M.-E. Long-Term Outcome of Patients with Relapsed/Refractory B-Cell Non-Hodgkin Lymphoma Treated with Blinatumomab. Blood Adv. 2019, 3, 2491–2498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. van der Horst, H.J.; de Jonge, A.V.; Hiemstra, I.H.; Gelderloos, A.T.; Berry, D.R.A.I.; Hijmering, N.J.; van Essen, H.F.; de Jong, D.; Chamuleau, M.E.D.; Zweegman, S.; et al. Epcoritamab Induces Potent Anti-Tumor Activity against Malignant B-Cells from Patients with DLBCL, FL and MCL, Irrespective of Prior CD20 Monoclonal Antibody Treatment. Blood Cancer J. 2021, 11, 38. [Google Scholar] [CrossRef] [PubMed]
  163. Van Der Horst, H.J.; de Jonge, A.V.; Hiemstra, I.H.; Gelderloos, A.T.; Berry, D.R.; Hijmering, N.J.; de Jong, D.; Chamuleau, M.E.D.; Zweegman, S.; Breij, E.C.; et al. Duobody-CD3xCD20 Induces Potent Anti-Tumor Activity in Malignant Lymph Node B Cells from Patients with DLBCL, FL and MCL Ex Vivo, Irrespective of Prior Treatment with CD20 Monoclonal Antibodies. Blood 2019, 134, 4066. [Google Scholar] [CrossRef]
  164. Hutchings, M.; Mous, R.; Clausen, M.R.; Johnson, P.; Linton, K.M.; Chamuleau, M.E.D.; Lewis, D.J.; Balari, A.S.; Cunningham, D.; Oliveri, R.S.; et al. Dose Escalation of Subcutaneous Epcoritamab in Patients with Relapsed or Refractory B-Cell Non-Hodgkin Lymphoma: An Open-Label, Phase 1/2 Study. Lancet 2021, 398, 1157–1169. [Google Scholar] [CrossRef]
  165. Glorius, P.; Baerenwaldt, A.; Kellner, C.; Staudinger, M.; Dechant, M.; Stauch, M.; Beurskens, F.J.; Parren, P.W.H.I.; van de Winkel, J.G.J.; Valerius, T.; et al. The Novel Tribody [(CD20)(2)XCD16] Efficiently Triggers Effector Cell-Mediated Lysis of Malignant B Cells. Leukemia 2013, 27, 190–201. [Google Scholar] [CrossRef] [Green Version]
  166. Chao, M.P.; Alizadeh, A.A.; Tang, C.; Myklebust, J.H.; Varghese, B.; Gill, S.; Jan, M.; Cha, A.C.; Chan, C.K.; Tan, B.T.; et al. Anti-CD47 Antibody Synergizes with Rituximab to Promote Phagocytosis and Eradicate Non-Hodgkin Lymphoma. Cell 2010, 142, 699–713. [Google Scholar] [CrossRef] [Green Version]
  167. Buatois, V.; Johnson, Z.; Salgado-Pires, S.; Papaioannou, A.; Hatterer, E.; Chauchet, X.; Richard, F.; Barba, L.; Daubeuf, B.; Cons, L.; et al. Preclinical Development of a Bispecific Antibody That Safely and Effectively Targets CD19 and CD47 for the Treatment of B-Cell Lymphoma and Leukemia. Mol. Cancer Ther. 2018, 17, 1739–1751. [Google Scholar] [CrossRef] [Green Version]
  168. Zeng, J.; Liu, R.; Wang, J.; Fang, Y. A Bispecific Antibody Directly Induces Lymphoma Cell Death by Simultaneously Targeting CD20 and HLA-DR. J. Cancer Res. Clin. Oncol. 2015, 141, 1899–1907. [Google Scholar] [CrossRef]
  169. Tuscano, J.M.; Ma, Y.; Martin, S.M.; Kato, J.; O’Donnell, R.T. The Bs20x22 Anti-CD20-CD22 Bispecific Antibody Has More Lymphomacidal Activity than Do the Parent Antibodies Alone. Cancer Immunol. Immunother. 2011, 60, 771–780. [Google Scholar] [CrossRef] [Green Version]
  170. Qu, Z.; Goldenberg, D.M.; Cardillo, T.M.; Shi, V.; Hansen, H.J.; Chang, C.-H. Bispecific Anti-CD20/22 Antibodies Inhibit B-Cell Lymphoma Proliferation by a Unique Mechanism of Action. Blood 2008, 111, 2211–2219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Senter, P.D. Potent Antibody Drug Conjugates for Cancer Therapy. Curr. Opin. Chem. Biol. 2009, 13, 235–244. [Google Scholar] [CrossRef]
  172. Gerber, H.-P.; Sapra, P.; Loganzo, F.; May, C. Combining Antibody–Drug Conjugates and Immune-Mediated Cancer Therapy: What to Expect? Biochem. Pharmacol. 2016, 102, 1–6. [Google Scholar] [CrossRef] [PubMed]
  173. Chu, Y.-W.; Polson, A. Antibody–Drug Conjugates for the Treatment of B-Cell Non-Hodgkin’s Lymphoma and Leukemia. Future Oncol. 2013, 9, 355–368. [Google Scholar] [CrossRef] [PubMed]
  174. Zinzani, P.L.; Sasse, S.; Radford, J.; Shonukan, O.; Bonthapally, V. Experience of Brentuximab Vedotin in Relapsed/Refractory Hodgkin Lymphoma and Relapsed/Refractory Systemic Anaplastic Large-Cell Lymphoma in the Named Patient Program: Review of the Literature. Crit. Rev. Oncol. Hematol. 2015, 95, 359–369. [Google Scholar] [CrossRef] [PubMed]
  175. Lamb, Y.N. Inotuzumab Ozogamicin: First Global Approval. Drugs 2017, 77, 1603–1610. [Google Scholar] [CrossRef] [PubMed]
  176. Sehn, L.H.; Herrera, A.F.; Flowers, C.R.; Kamdar, M.K.; McMillan, A.; Hertzberg, M.; Assouline, S.; Kim, T.M.; Kim, W.S.; Ozcan, M.; et al. Polatuzumab Vedotin in Relapsed or Refractory Diffuse Large B-Cell Lymphoma. J. Clin. Oncol. 2020, 38, 155–165. [Google Scholar] [CrossRef]
  177. Loganzo, F.; Sung, M.; Gerber, H.-P. Mechanisms of Resistance to Antibody–Drug Conjugates. Mol. Cancer Ther. 2016, 15, 2825–2834. [Google Scholar] [CrossRef] [Green Version]
  178. Drago, J.Z.; Modi, S.; Chandarlapaty, S. Unlocking the Potential of Antibody–Drug Conjugates for Cancer Therapy. Nat. Rev. Clin. Oncol. 2021, 18, 327–344. [Google Scholar] [CrossRef] [PubMed]
  179. Szakács, G.; Paterson, J.K.; Ludwig, J.A.; Booth-Genthe, C.; Gottesman, M.M. Targeting Multidrug Resistance in Cancer. Nat. Rev. Drug Discov. 2006, 5, 219–234. [Google Scholar] [CrossRef]
  180. Chen, R.; Herrera, A.F.; Hou, J.; Chen, L.; Wu, J.; Guo, Y.; Synold, T.W.; Ngo, V.N.; Puverel, S.; Mei, M.; et al. Inhibition of MDR1 Overcomes Resistance to Brentuximab Vedotin in Hodgkin Lymphoma. Clin. Cancer Res. 2020, 26, 1034–1044. [Google Scholar] [CrossRef]
  181. Qing, K.; Jin, Z.; Xu, Z.; Wang, W.; Li, X.; Zhang, Y.; Wang, L.; Zhu, H.; Xiang, R.; Wu, S.; et al. Dysregulated MDR1 by PRDM1/Blimp1 Is Involved in the Doxorubicin Resistance of Non-Germinal Center B-Cell-Like Diffuse Large B-Cell Lymphoma. Chemotherapy 2021, 1–12. [Google Scholar] [CrossRef]
  182. Beck, A.; Goetsch, L.; Dumontet, C.; Corvaïa, N. Strategies and Challenges for the next Generation of Antibody–Drug Conjugates. Nat. Rev. Drug Discov. 2017, 16, 315–337. [Google Scholar] [CrossRef] [PubMed]
  183. Buecheler, J.W.; Winzer, M.; Weber, C.; Gieseler, H. Alteration of Physicochemical Properties for Antibody-Drug Conjugates and Their Impact on. Stability. J. Pharm. Sci. 2020, 109, 161–168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Collins, D.M.; Bossenmaier, B.; Kollmorgen, G.; Niederfellner, G. Acquired Resistance to Antibody-Drug Conjugates. Cancers 2019, 11, 394. [Google Scholar] [CrossRef] [Green Version]
  185. Vaklavas, C.; Forero-Torres, A. Safety and Efficacy of Brentuximab Vedotin in Patients with Hodgkin Lymphoma or Systemic Anaplastic Large Cell Lymphoma. Ther. Adv. Hematol. 2012, 3, 209–225. [Google Scholar] [CrossRef] [Green Version]
  186. DiJoseph, J.F.; Dougher, M.M.; Kalyandrug, L.B.; Armellino, D.C.; Boghaert, E.R.; Hamann, P.R.; Moran, J.K.; Damle, N.K. Antitumor Efficacy of a Combination of CMC-544 (Inotuzumab Ozogamicin), a CD22-Targeted Cytotoxic Immunoconjugate of Calicheamicin, and Rituximab against Non-Hodgkin’s B-Cell Lymphoma. Clin. Cancer Res. 2006, 12, 242–249. [Google Scholar] [CrossRef] [Green Version]
  187. Sawas, A.; Savage, K.J.; Perez, R.P.; Advani, R.H.; Melhem-Bertrandt, A.; Lackey, J.; Trave, F.; Anand, B.; Huang, Y.; Vincent, M.; et al. A First in Human Experience of the Anti-CD37 Antibody-Drug Conjugate AGS67E in Lymphoid Malignancies. J. Clin. Oncol. 2016, 34, 7549. [Google Scholar] [CrossRef]
  188. Advani, R.H.; Lebovic, D.; Chen, A.; Brunvand, M.; Goy, A.; Chang, J.E.; Hochberg, E.; Yalamanchili, S.; Kahn, R.; Lu, D.; et al. Phase I Study of the Anti-CD22 Antibody–Drug Conjugate Pinatuzumab Vedotin with/without Rituximab in Patients with Relapsed/Refractory B-Cell Non-Hodgkin Lymphoma. Clin. Cancer Res. 2017, 23, 1167–1176. [Google Scholar] [CrossRef] [Green Version]
  189. Stathis, A.; Freedman, A.S.; Flinn, I.W.; Maddocks, K.J.; Weitman, S.; Berdeja, J.G.; Mejia, A.V.; Zucca, E.; Green, R.; Romanelli, A.; et al. A Phase I Study of IMGN529, an Antibody-Drug Conjugate (ADC) Targeting CD37, in Adult Patients with Relapsed or Refractory B-Cell Non-Hodgkin’s Lymphoma (NHL). Blood 2014, 124, 1760. [Google Scholar] [CrossRef]
  190. Fathi, A.T.; Borate, U.; DeAngelo, D.J.; O’Brien, M.M.; Trippett, T.; Shah, B.D.; Hale, G.A.; Foran, J.M.; Silverman, L.B.; Tibes, R.; et al. A Phase 1 Study of Denintuzumab Mafodotin (SGN-CD19A) in Adults with Relapsed or Refractory B-Lineage Acute Leukemia (B-ALL) and Highly Aggressive Lymphoma. Blood 2015, 126, 1328. [Google Scholar] [CrossRef]
  191. Trnĕný, M.; Verhoef, G.; Dyer, M.J.; Ben Yehuda, D.; Patti, C.; Canales, M.; Lopez, A.; Awan, F.T.; Montgomery, P.G.; Janikova, A.; et al. A Phase II Multicenter Study of the Anti-CD19 Antibody Drug Conjugate Coltuximab Ravtansine (SAR3419) in Patients with Relapsed or Refractory Diffuse Large B-Cell Lymphoma Previously Treated with Rituximab-Based Immunotherapy. Haematologica 2018, 103, 1351–1358. [Google Scholar] [CrossRef] [PubMed]
  192. Zhylko, A.; Winiarska, M.; Graczyk-Jarzynka, A. The Great War of Today: Modifications of CAR-T Cells to Effectively Combat Malignancies. Cancers 2020, 12, 2030. [Google Scholar] [CrossRef] [PubMed]
  193. Miazek-Zapala, N.; Slusarczyk, A.; Kusowska, A.; Zapala, P.; Kubacz, M.; Winiarska, M.; Bobrowicz, M. The “Magic Bullet” Is Here? Cell-Based Immunotherapies for Hematological Malignancies in the Twilight of the Chemotherapy Era. Cells 2021, 10, 1511. [Google Scholar] [CrossRef] [PubMed]
  194. Prasad, V. Immunotherapy: Tisagenlecleucel-the First Approved CAR-T-Cell Therapy: Implications for Payers and Policy Makers. Nat. Rev. Clin. Oncol. 2018, 15, 11–12. [Google Scholar] [CrossRef]
  195. Al-Mansour, M.; Al-Foheidi, M.; Ibrahim, E. Efficacy and Safety of Second-Generation CAR T-Cell Therapy in Diffuse Large B-Cell Lymphoma: A Meta-Analysis. Mol. Clin. Oncol. 2020, 13, 33. [Google Scholar] [CrossRef]
  196. Maloney, D.G. Anti-CD19 CAR T Cell Therapy for Lymphoma-off to the Races! Nat. Rev. Clin. Oncol. 2019, 16, 279–280. [Google Scholar] [CrossRef]
  197. Shah, N.N.; Fry, T.J. Mechanisms of Resistance to CAR T Cell Therapy. Nat. Rev. Clin. Oncol. 2019, 16, 372–385. [Google Scholar] [CrossRef]
  198. Laurent, C.; Syrykh, C.; Hamon, M.; Adélaïde, J.; Guille, A.; Escudié, F.; Jalowicki, G.; Fina, F.; Bardet, A.; Mescam, L.; et al. Resistance of B-Cell Lymphomas to CAR T-Cell Therapy Is Associated With Genomic Tumor Changes Which Can Result in Transdifferentiation. Am. J. Surg. Pathol. 2021. [Google Scholar] [CrossRef]
  199. Majzner, R.G.; Mackall, C.L. Tumor Antigen Escape from CAR T-Cell Therapy. Cancer Discov. 2018, 8, 1219–1226. [Google Scholar] [CrossRef] [Green Version]
  200. Ruella, M.; Maus, M.V. Catch Me If You Can: Leukemia Escape after CD19-Directed T Cell Immunotherapies. Comput. Struct. Biotechnol. J. 2016, 14, 357–362. [Google Scholar] [CrossRef] [Green Version]
  201. Gardner, R.; Wu, D.; Cherian, S.; Fang, M.; Hanafi, L.-A.; Finney, O.; Smithers, H.; Jensen, M.C.; Riddell, S.R.; Maloney, D.G.; et al. Acquisition of a CD19-Negative Myeloid Phenotype Allows Immune Escape of MLL-Rearranged B-ALL from CD19 CAR-T-Cell Therapy. Blood 2016, 127, 2406–2410. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Sotillo, E.; Barrett, D.M.; Black, K.L.; Bagashev, A.; Oldridge, D.; Wu, G.; Sussman, R.; Lanauze, C.; Ruella, M.; Gazzara, M.R.; et al. Convergence of Acquired Mutations and Alternative Splicing of CD19 Enables Resistance to CART-19 Immunotherapy. Cancer Discov. 2015, 5, 1282–1295. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Ledererova, A.; Dostalova, L.; Kozlova, V.; Peschelova, H.; Ladungova, A.; Culen, M.; Loja, T.; Verner, J.; Pospisilova, S.; Smida, M.; et al. Hypermethylation of CD19 Promoter Enables Antigen-Negative Escape to CART-19 in Vivo and in Vitro. J. Immunother. Cancer 2021, 9, e002352. [Google Scholar] [CrossRef] [PubMed]
  204. Hamieh, M.; Dobrin, A.; Cabriolu, A.; van der Stegen, S.J.C.; Giavridis, T.; Mansilla-Soto, J.; Eyquem, J.; Zhao, Z.; Whitlock, B.M.; Miele, M.M.; et al. CAR T Cell Trogocytosis and Cooperative Killing Regulate Tumour Antigen Escape. Nature 2019, 568, 112–116. [Google Scholar] [CrossRef] [PubMed]
  205. Jain, M.D.; Ziccheddu, B.; Coughlin, C.A.; Faramand, R.; Griswold, A.J.; Reid, K.M.; Landgren, O.; Locke, F.L.; Maura, F.; Davila, M.L.; et al. Genomic Drivers of Large B-Cell Lymphoma Resistance to CD19 CAR-T Therapy. Blood 2021, 138, 42. [Google Scholar] [CrossRef]
  206. Fousek, K.; Watanabe, J.; Joseph, S.K.; George, A.; An, X.; Byrd, T.T.; Morris, J.S.; Luong, A.; Martínez-Paniagua, M.A.; Sanber, K.; et al. CAR T-Cells That Target Acute B-Lineage Leukemia Irrespective of CD19 Expression. Leukemia 2021, 35, 75–89. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Shalabi, H.; Kraft, I.L.; Wang, H.-W.; Yuan, C.M.; Yates, B.; Delbrook, C.; Zimbelman, J.D.; Giller, R.; Stetler-Stevenson, M.; Jaffe, E.S.; et al. Sequential Loss of Tumor Surface Antigens Following Chimeric Antigen Receptor T-Cell Therapies in Diffuse Large B-Cell Lymphoma. Haematologica 2018, 103, e215–e218. [Google Scholar] [CrossRef]
  208. Majzner, R.G.; Frank, J.M.; Mount, C.; Tousley, A.; Kurtz, D.M.; Sworder, B.; Murphy, K.A.; Manousopoulou, A.; Kohler, K.; Rotiroti, M.C.; et al. CD58 Aberrations Limit Durable Responses to CD19 CAR in Large B Cell Lymphoma Patients Treated with Axicabtagene Ciloleucel but Can Be Overcome through Novel CAR Engineering. Blood 2020, 136, 53. [Google Scholar] [CrossRef]
  209. Singh, N.; Lee, Y.G.; Shestova, O.; Ravikumar, P.; Hayer, K.E.; Hong, S.J.; Lu, X.M.; Pajarillo, R.; Agarwal, S.; Kuramitsu, S.; et al. Impaired Death Receptor Signaling in Leukemia Causes Antigen-Independent Resistance by Inducing CAR T-Cell Dysfunction. Cancer Discov. 2020, 10, 552–567. [Google Scholar] [CrossRef] [Green Version]
  210. Cai, Q.; Zhang, M.; Li, Z. Potential Strategies against Resistance to CAR T-Cell Therapy in Haematological Malignancies. Ther. Adv. Med. Oncol. 2020, 12, 1758835920962963. [Google Scholar] [CrossRef]
  211. Liu, Y.; Barta, S.K. Diffuse Large B-Cell Lymphoma: 2019 Update on Diagnosis, Risk Stratification, and Treatment. Am. J. Hematol. 2019, 94, 604–616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Mechanisms of resistance to anti-CD20-based immunotherapies in DLBCL. (A) Induction of resistance to complement-dependent cytotoxicity (CDC) is mediated by an increased expression of complement regulatory proteins (CRPs)–CD46, CD55, CD59 and soluble protective protein complement factor H (CFH). The abnormal composition of the lipid raft domains, e.g., hypersialylation compromises complement-induced cell lysis. (B) Impaired antibody-dependent cellular cytotoxicity (ADCC) is caused by genetic polymorphism in FcγRIII receptor (CD16) and low expression of CD16. Interactions between specific KIR ligands on NK cells and HLA molecules on B cells reduce NK-cell degranulation. Monocytes mediate shaving of RTX/CD20 complexes from the cell surface; macrophages and neutrophils reduce the anti-tumor potential of NK cells by releasing reactive oxygen species (ROS). (C) CD20 expression is regulated by numerous epigenetic and transcription factors, e.g., Foxo1. Deficiency in CD20 level arises from mutations in MS4A1 gene leading to alternative splicing, and from CD20 internalization. Conformational changes in CD20 protein decrease binding affinity to anti-CD20 mAbs.
Figure 1. Mechanisms of resistance to anti-CD20-based immunotherapies in DLBCL. (A) Induction of resistance to complement-dependent cytotoxicity (CDC) is mediated by an increased expression of complement regulatory proteins (CRPs)–CD46, CD55, CD59 and soluble protective protein complement factor H (CFH). The abnormal composition of the lipid raft domains, e.g., hypersialylation compromises complement-induced cell lysis. (B) Impaired antibody-dependent cellular cytotoxicity (ADCC) is caused by genetic polymorphism in FcγRIII receptor (CD16) and low expression of CD16. Interactions between specific KIR ligands on NK cells and HLA molecules on B cells reduce NK-cell degranulation. Monocytes mediate shaving of RTX/CD20 complexes from the cell surface; macrophages and neutrophils reduce the anti-tumor potential of NK cells by releasing reactive oxygen species (ROS). (C) CD20 expression is regulated by numerous epigenetic and transcription factors, e.g., Foxo1. Deficiency in CD20 level arises from mutations in MS4A1 gene leading to alternative splicing, and from CD20 internalization. Conformational changes in CD20 protein decrease binding affinity to anti-CD20 mAbs.
Ijms 23 01501 g001
Figure 2. Targeted immunotherapies in hematological malignancies. (A) bispecific antibody; (B) trispecific antibody; (C) antibody–drug conjugate; (D) chimeric antigen receptor T cells.
Figure 2. Targeted immunotherapies in hematological malignancies. (A) bispecific antibody; (B) trispecific antibody; (C) antibody–drug conjugate; (D) chimeric antigen receptor T cells.
Ijms 23 01501 g002
Table 1. Selected bispecific antibodies in clinical and preclinical studies.
Table 1. Selected bispecific antibodies in clinical and preclinical studies.
NameTargetTypeMode of ActionEfficacy in Clinical and Preclinical Studies
BlinatumomabCD3/CD19bispecific antibody (BiTE)CD3-positive T-cell recruitment
  • In phase 2 study on r/r DLBCL patients (NCT01741792), who received 3 prior lines of therapy blinatumomab induced OR in 43% of patients (9/21), including CR in 19% of subjects (4/21) [159].
  • In heavily pre-treated r/r B-NHL patients (n = 73) receiving blinatumomab long-term remission (over 7 years) was achieved in a subset of patients [160,161].
Epcoritamab
(DuoBody-CD3xCD20)
CD3/CD20bispecific antibody (BiTE)CD3-positive T-cell recruitment
  • Epcoritamab induced potent cytotoxicity against primary B-NHL samples (DLBCL, FL, MCL) irrespective of whether the patients were newly diagnosed or were r/r after CD20-mAb treatment [162,163].
  • In phase 1/2 clinical trial on 68 r/r B-NHL patients (NCT03625037) epcoritamab showed ORR in 86% of patients, including CR in 45% of subjects with no grade 3 or higher adverse events [164].
[(CD20)2xCD16]CD20/CD16bispecific tribodyCD16-positive effector cell recruitment
  • In preclinical studies ((CD20)2xCD16) demonstrated a 9-fold enhancement in ADCC and promoted 10-fold higher NK-cell number activation in comparison to RTX, irrespective of FcγRIIIa polymorphism [165].
NI-1701CD19/CD47bispecific antibodyADCP enhancement via blocking immune checkpoint receptor
  • Blocking CD47 with mAb increased phagocytosis of NHL cells in preclinical studies [166] and in a small clinical study in DLBCL patients (NCT02953509).
  • NI-1701 induced potent cytotoxicity against B-lymphoma and leukemic cell lines in vitro and in NOD/SCID mouse xenograft model. It was more effective than monovalent anti-CD19 and anti-CD47 mAbs and achieved a synergistic effect with RTX [167].
CD20-HLA-DR DVD-IgCD20/HLA-DRbispecific antibodyIncrease in selectivity against NHL cells versus healthy cells
  • CD20-HLA-DR DVD-Ig demonstrated potent CDC and ADCC against B-cell lymphoma in vitro by inducing homotypic adhesion and actin reorganization. It effectively depleted Raji cells from a mixture with whole human blood and showed high specificity towards malignant cells, leaving normal B cells unaffected [168].
Bs20×22CD20/CD22bispecific antibody (RTX/HB22.7 platform)Induction of apoptosis
  • Bs20x22 induced a 3-fold higher rate of apoptosis than parent mAbs (anti-CD20 RTX and anti-CD22 HB22.7) despite unchanged binding affinity. It also increased survival in nude mice Raji xenograft model (88%) as compared to other treatment regimens (RTX and Bs20x22–75%, RTX-50%, HB22.7–25%) [169].
Bs20x22CD20/CD22bispecific antibody (hA20 (veltuzumab)/hLL2 (epratuzumab) platform)ADCC enhancement via antigen crosslinking
  • Tetravalent anti-CD20/22 bsAb constructs demonstrated remarkable redistribution of CD20, CD22, and BCR into lipid rafts following cross-linking of both antigens leading to the initiation of phosphorylation cascade and internalization of Ab-BCR complexes [170].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kusowska, A.; Kubacz, M.; Krawczyk, M.; Slusarczyk, A.; Winiarska, M.; Bobrowicz, M. Molecular Aspects of Resistance to Immunotherapies—Advances in Understanding and Management of Diffuse Large B-Cell Lymphoma. Int. J. Mol. Sci. 2022, 23, 1501. https://doi.org/10.3390/ijms23031501

AMA Style

Kusowska A, Kubacz M, Krawczyk M, Slusarczyk A, Winiarska M, Bobrowicz M. Molecular Aspects of Resistance to Immunotherapies—Advances in Understanding and Management of Diffuse Large B-Cell Lymphoma. International Journal of Molecular Sciences. 2022; 23(3):1501. https://doi.org/10.3390/ijms23031501

Chicago/Turabian Style

Kusowska, Aleksandra, Matylda Kubacz, Marta Krawczyk, Aleksander Slusarczyk, Magdalena Winiarska, and Malgorzata Bobrowicz. 2022. "Molecular Aspects of Resistance to Immunotherapies—Advances in Understanding and Management of Diffuse Large B-Cell Lymphoma" International Journal of Molecular Sciences 23, no. 3: 1501. https://doi.org/10.3390/ijms23031501

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