Next Article in Journal
Assessment of Ovarian Tumor Growth in Wild-Type and Lumican-Deficient Mice: Insights Using Infrared Spectral Imaging, Histopathology, and Immunohistochemistry
Next Article in Special Issue
Clinical Studies on Cytokine-Induced Killer Cells: Lessons from Lymphoma Trials
Previous Article in Journal
Effect of Adjuvant and Palliative Chemotherapy in Large Cell Neuroendocrine Carcinoma of the Lung: A Systematic Review and Meta-Analysis
Previous Article in Special Issue
Acceptability and Preliminary Efficacy of a Web- and Telephone-Based Personalised Exercise Intervention for Individuals with Metastatic Prostate Cancer: The ExerciseGuide Pilot Randomised Controlled Trial
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Genome Instability in Multiple Myeloma: Facts and Factors

1
Laboratory of Amyloid Biology, St. Petersburg State University, 199034 St. Petersburg, Russia
2
International Laboratory “Computer Technologies”, ITMO University, 197101 St. Petersburg, Russia
3
Department of Microbiology and Molecular Genetics, University of California, Davis, CA 95616, USA
4
Department of Genetics and Biotechnology, St. Petersburg State University, 199034 St. Petersburg, Russia
5
Vavilov Institute of General Genetics, St. Petersburg Branch, Russian Academy of Sciences, 199034 St. Petersburg, Russia
6
Russian Research Institute of Hematology and Transfusiology, 191024 St. Petersburg, Russia
7
Eppley Institute for Research in Cancer, Fred and Pamela Buffett Cancer Center, University of Nebraska Medical Center, Omaha, NE 68198, USA
8
Departments of Biochemistry and Molecular Biology, Microbiology and Pathology, Genetics Cell Biology and Anatomy, University of Nebraska Medical Center, Omaha, NE 68198, USA
*
Authors to whom correspondence should be addressed.
Cancers 2021, 13(23), 5949; https://doi.org/10.3390/cancers13235949
Submission received: 31 August 2021 / Revised: 20 October 2021 / Accepted: 22 November 2021 / Published: 26 November 2021
(This article belongs to the Special Issue Feature Paper from Journal Reviewers)

Abstract

:

Simple Summary

Multiple myeloma is an incurable blood cancer caused by the malignant transformation of immunoglobulin-producing plasma cells. The mechanisms leading to the origin of cancerous cells and the evolution of myeloma disease are not understood. The development of myeloma is accompanied by genetic changes affecting various cellular pathways. This review describes current progress in understanding the etiology of the disease that might stimulate the development of new therapies.

Abstract

Multiple myeloma (MM) is a malignant neoplasm of terminally differentiated immunoglobulin-producing B lymphocytes called plasma cells. MM is the second most common hematologic malignancy, and it poses a heavy economic and social burden because it remains incurable and confers a profound disability to patients. Despite current progress in MM treatment, the disease invariably recurs, even after the transplantation of autologous hematopoietic stem cells (ASCT). Biological processes leading to a pathological myeloma clone and the mechanisms of further evolution of the disease are far from complete understanding. Genetically, MM is a complex disease that demonstrates a high level of heterogeneity. Myeloma genomes carry numerous genetic changes, including structural genome variations and chromosomal gains and losses, and these changes occur in combinations with point mutations affecting various cellular pathways, including genome maintenance. MM genome instability in its extreme is manifested in mutation kataegis and complex genomic rearrangements: chromothripsis, templated insertions, and chromoplexy. Chemotherapeutic agents used to treat MM add another level of complexity because many of them exacerbate genome instability. Genome abnormalities are driver events and deciphering their mechanisms will help understand the causes of MM and play a pivotal role in developing new therapies.

1. Clinical Manifestation of Multiple Myeloma and Recent Research Approaches

1.1. Clinical Characteristics of Pre-MM and MM: Heterogeneity and Clonal Evolution of Cancer Cells

MM is characterized by aberrant expansion of terminally differentiated monoclonal plasma cells resulting in symptoms described by the acronym “CRAB”: hypercalcemia, renal failure, anemia, and bone lesions. In addition to that, diagnostic criteria include three biomarkers of malignancy: the presence of excessive clonal bone marrow plasma cells, elevated serum free light chain ratio (ratio of κ to λ free light chains), or focal bone lesions [1,2,3] (Figure 1). These symptoms and biomarkers are called myeloma-defining events (MDE) [1]. At least one MDE in addition to a biopsy-proven plasmacytoma or ≥10% of plasma cells in bone marrow is required for the MM diagnosis [1]. Development of MM is a multi-stage process beginning from a premalignant stage termed monoclonal gammopathy of undetermined significance (MGUS) [4,5,6]. MGUS can be accidentally found years to decades before the actual diagnosis of MM. It could be present in ~3% of the normal human population over 50 years old [7]. MGUS does not necessarily develop into MM; further progression of MGUS into the active MM has linear risk and occurs with a rate of ~1% per year [1,8,9,10,11].
An additional, more advanced stage is observed in some patients, which is referred to as smoldering multiple myeloma (SMM) [1,11,12,13,14]. MGUS and SMM are usually asymptomatic stages characterized by different levels of M-protein production and different ratios of clonal plasma cells in bone marrow (see Figure 1). SMM may represent asymptomatic MM rather than being an MM precursor [15]. Patients with SMM follow a declining logarithmic progression curve to symptomatic MM: 10% risk per year for the first 5 years following diagnosis, 3% risk per year for the following 5 years, and a subsequent 1% risk per year [1,8,13]. Both MGUS and SMM have specific diagnostic criteria (Figure 1). MM and its precursors can produce different types of monoclonal proteins. For instance, the production of different isotypes of the immunoglobulin heavy chain has been described in MGUS [7]. The two clinically significant entities are IgM MGUS and non-IgM MGUS demonstrating different chances of progression into the MM. While the non-IgM MGUS is associated with a risk of progression to MM, IgM MGUS most frequently progresses into non-Hodgkin lymphoma and its subtype Waldenström macroglobulinemia [10,16,17]. In addition, no immunoglobulin heavy chain production can be seen in some cases in patients with abnormal serum free light chain ratio, which is attributed to the light chain MGUS and the light chain MM [6,18]. An uncommon subtype of MM is non-secretory myeloma, found in 2–3% of all MM cases [1,18,19].
The biological processes leading to the appearance of a pathological myeloma clone and the mechanisms of the disease evolution are not yet deciphered. Factors responsible for the emergence of MM include a combination of genetic predisposition, alterations in genomes of the lymphoid cells (such as IGH translocations or hyperdiploidy), and a variety of secondary changes, which include accumulation of mutations, chromosomal rearrangements, and complex genetic events (Figure 1). Apparent inefficiency of the initial treatment, fast recurrences, resistance to the earlier prescribed drugs, and clonal behavior of the disease implies that MM represents a heterogeneous entity. Multiple studies suggest that newly diagnosed MM represents an aggregate of the main pathological clone with several subclones that acquire proliferative priority when the main clone is suppressed during specific therapy [20,21,22,23,24,25,26,27,28,29,30,31,32,33,34].
The classic clonal evolution implies the sequential acquisition of mutations with a concomitant sequential selection of successive subclones, their expansion, and mutual interference [35,36]. Each tumor cell can carry many genetic abnormalities, including mutations that provide a selective growth advantage: “driver” mutations, selectively neutral “passenger” mutations, and deleterious mutations affecting fitness [37]. In addition to that, there are “mutator” mutations that increase the rate of genetic changes. The dynamics of tumor evolution is a function of mutation rate elevation and clonal expansion that relies on “driver” mutations. Natural selection provides “selective sweeps” when one or several clones grow to dominate the neoplasm [36]. These dominant clones accumulate new genetic changes in addition to the mutational landscape of the original tumor as the evolution proceeds. The complexity of this process is augmented by epigenetic changes that, similar to changes in the DNA, can confer either growth advantage or be selectively neutral or deleterious. Epigenetic changes can also affect mutation rates. On top of this complexity is the impact of therapeutic drugs, which can modify DNA, affect DNA repair processes, and modulate the growth advantage of the cancer cells [38,39,40].
The clonal development of the tumor may follow the linear model if driver changes provide a strong selective advantage that outcompetes all previous clones, be branched if several clones expand simultaneously, or neutral if there is no selective advantage between multiple clones’ appearance and co-existence. Alternatively, most essential changes may occur simultaneously or near-simultaneously early in tumor development, establishing several dominant clones that grow stably, a characteristic of punctuated evolution [36,41]. It is generally recognized that the MM development follows the rules of branched evolution [27,42,43]. However, patterns of clones’ phylogeny consistent with linear, neutral, and punctuated models of evolution have also been described during MM development and progression [22,27,29,44,45]. Moreover, recent advances in studies of MM and its precursors have centered the punctuated evolution model as characterizing best the early stages of the MM development [20,46,47,48,49]. For instance, significant heterogeneity is present before the disease manifestation at the MGUS and SMM stages [20,23,30,47,48,50,51,52], and further progression from MGUS/SMM to MM is characterized by clonal stability in some tumors [20,46,47,48]. Importantly, clonal evolution in MM seems to be closely interrelated with treatment strategies [22,23,24,25,26,28,34,45,53,54]. As the disease proceeds, the variability of the clones increases, which creates the ground for further clonal diversification and evolution [23,45,47,51,52,53,55,56,57,58,59,60,61,62,63]. One or several such clones will ultimately lead to the recurrence of myeloma. It is possible that the change from the main clone to subclones during MM treatment causes a change in the clinical and hematological phenotype of MM and, thus, explains the inefficiency of the previously conducted therapy. Noteworthy, patients with MM or its precursors have an increased risk of developing secondary primary malignancies (SPMs), such as myelodysplastic syndrome (MDS), acute myeloid leukemia (AML), and others [64,65,66,67]. The origin of these tumors is likely related to the genotoxic action of some therapeutic agents [65,68,69,70,71,72,73,74], although an excess risk for hematopoietic neoplasms other than MM in MGUS patients supports an idea that endogenous factors play an essential role in SPMs’ development [9,75,76,77]. Therefore, it is possible that the genetic landscape of hematopoietic stem cells in MM patients may predispose them to different malignant programs that can unfold spontaneously or as a result of therapeutic intervention.
Recent developments and advances of next-generation sequencing technologies (NGS) help to produce a large amount of genomic data that are invaluable for MM diagnosis, choices of possible treatment, assessment of drug response, and understanding of disease evolution [78,79,80]. During the last decade, dozens of large-scale NGS studies have been undertaken to find peculiarities of the MM genomes and pinpoint genetic drivers of MM (Table S1) [28,29,61,81,82,83]. Whole-exome (WES) or whole-genome (WGS) sequencing of MM genomes allows for detecting point mutations, small insertions or deletions, and structural variations. Many studies combine their own NGS data with the publicly available datasets from Multiple Myeloma Research Foundation (MMRF) CoMMpass Study (Clinicaltrials.gov identification number: NCT01454297) [84]. Sequencing data for MM samples are deposited in the European Genome Archive or at the Genotype and Phenotype database (dbGaP). The Multiple Myeloma Genome Project (MGP) attempts to assemble and analyze NGS data for MM, improve clinical testing, and define treatment strategies for MM patients [85,86]. MGP provides the repository of WES, WGS, and RNA-Seq data for patients with MM derived from different sources, such as the Multiple Myeloma Research Foundation, the Myeloma XI trial, and others [85].

1.2. Current Treatment Algorithms of MM: Mechanisms of Action of Anti-MM Agents

MM patient management strategy differs significantly depending on age, comorbidities, cytogenetic parameters, disease stage, risk stratification, and other factors [87,88,89,90,91] (Figure 2). For the treatment of primary patients under 70 years of age without serious comorbidities, high-dose chemotherapy (HDCT) with autologous hematopoietic stem cell transplantation (ASCT) is included in the treatment program (Figure 2). Several recent studies show that ASCT can be a safe choice for patients older than 70 years [92,93,94,95,96]. Despite significant toxicity, ASCT remains the standard of care and is associated with prolonged progression-free survival and better overall survival [87,88,90,97,98,99,100,101,102,103,104,105,106,107,108,109]. ASCT can be delayed in standard-risk patients if they respond to the induction treatment [87,110]. HDCT with concomitant ASCT is not recommended for patients with comorbidities and for frail patients [87,88]. Examples of treating schemes for ASCT-ineligible patients are shown in Figure 2.
The therapeutic arsenal for MM treatment includes DNA damaging agents (melphalan, cyclophosphamide, etc.), immunomodulatory agents (IMiDs, e.g., thalidomide, lenalidomide, pomalidomide), proteasome inhibitors (PIs, e.g., bortezomib, carfilzomib, ixazomib), monoclonal antibodies (daratumumab, isatuximab, elotuzumab), and corticosteroids [1,87,91]. In addition to that, several drugs can be used in combination with other therapies as they are not considered efficient alone in most cases. These include inhibitors of histone deacetylase 6 (panobinostat), an inhibitor of Exportin-1 (selinexor, considered efficient in combination with dexamethazone), the DNA intercalating drugs anthracyclines (doxorubicin), and the BCL2 inhibitor venetoclax, which does not yet have approval for treatment of MM but appears to be efficient against MM with t(11;14) rearrangement [87,126,127]. Besides that, new agents and therapies have been approved or are about to be approved for the treatment of MM. They include an immunoconjugate drug targeting B-cell maturation antigen, BCMA (Belantamab Mafodotin), an inhibitor of kinesin spindle protein, KSP (Filanesib), chimeric antigen receptor (CAR) T-cell therapy (e.g., bb2121 therapy that targets BCMA), and bispecific antibodies [128,129,130,131,132,133,134].
Many agents used in MM treatment are genotoxic and can further elevate the genetic variability of MM, which can be harmful and potentiate disease progression and relapse [25,135]. The consequences of using one or another agent for the genome differ depending on the preexisting genomic alterations and mutations. Drugs within the chemotherapeutic arsenal with potential genotoxicity include alkylating DNA agents (melphalan, cyclophosphamide, bendamustine, busulfan), intercalating DNA agents (doxorubicin), microtubule-depolymerizing drugs (vincristine), agents inactivating topoisomerase II (etoposide, doxorubucin), and crosslinking agents (cisplatin). Melphalan is a phenylalanine-substituted derivative of nitrogen mustard that alkylates adenine and guanine in the DNA. It has two highly reactive chloroethylamine groups and thus induces intra- or intermolecular crosslinks in the DNA and between DNA and proteins [136,137,138]. The high-dose melphalan is a standard regimen before ASCT [87,103,136,139,140]. Cyclophosphamide also belongs to a family of mustard-alkylating agents and induces alkylation and crosslinking of the DNA [141,142,143]. At low doses, cyclophosphamide demonstrates immunomodulatory activity [144]. Notably, a significant increase in mutation burden and specific mutational signature have been reported in MM patients exposed to high-dose melphalan [54,59,63,145,146]. Melphalan (and to a lesser extent cyclophosphamide) is also associated with an increased risk of SPMs in MM patients [64,68,69,71,72,147,148,149,150]. The line of chemotherapeutic drugs has been consistently elaborated upon with novel therapeutic agents in recent years. One of them, a peptide-conjugate alkylator drug, melphalan flufenamide (melflufen), is currently in clinical trials [151,152,153]. Alkylating and crosslinking agents induce modifications to DNA that are a subject of DNA repair to allow for DNA replication and transcription [154]. The vulnerability of cancer cells to these drugs depends on the competing processes of cell cycle progression and repair. Base excision repair (BER), nucleotide excision repair (NER), and mismatch repair (MMR) remove monoadducts in the DNA, intra-strand crosslinks, and other lesions that affect only one strand of the two strands of the duplex DNA [155]. The DNA double-strand breaks (DSBs) interrupt the continuity of the DNA molecule and are repaired via two major repair routes: homologous recombination repair (HRR) and nonhomologous end-joining (NHEJ) [155]. The first one uses the homologous sequence as an instruction to recover the lost information, and the second one just joins the two ends together and seals the break. Despite their provisory division into error-free and error-prone pathways, correspondingly, both HRR and NHEJ can lead to point mutations and chromosome rearrangements. Inter-strand crosslinks constitute a physical barrier to the progression of both RNA and DNA polymerases since they impede the unwinding of the double helix. This type of lesion requires the action of multiple players belonging to the BER, NER, MMR, and HRR pathways and are grouped into a specialized FA/BRCA pathway [156,157]. FA/BRCA was named after Fanconi anemia (FA), a genetic disorder leading to developmental abnormalities and predisposition to cancer development [158]. Mutations in the FA/BRCA pathway genes make cells hypersensitive to agents inducing DNA crosslinks [156,159]. Upregulation of the FA/BRCA pathway contributes to acquired resistance to melphalan in MM cell lines [160]. Human myeloma cell lines frequently acquire mutations in the genes of the FA/BRCA pathway [161]. Selinexor, an inhibitor of the nuclear Exportin 1 (XPO1), decreases the expression of the FA/BRCA and NF-kB pathway genes, reduces melphalan-induced monoubiquitination of FANCD2, and overcomes the resistance of the MM cell lines to melphalan [162]. In addition to that, several other key factors of HR and NHEJ are found upregulated in MM cell lines, which can have implications to MM resistance to DNA-damaging drugs [163,164,165,166]. Bendamustine is a derivative of mechlorethamine that acts both as an alkylating agent and purine analog [167,168]. It has anti-MM activity and can overcome resistance to melphalan in MM cell lines [169]. It is most often used for the treatment of relapsed or refractory-to-other-regimens MM and patients with renal impairment not eligible for ASCT but it can be used as a conditioning therapy before ASCT when combined with melphalan [140,170,171,172,173,174,175,176]. Bendamustine efficiently activates the DNA damage response and has a synergistic effect with alkylating agents and pyrimidine analogs in killing MM cell lines [168].
IMiDs are derivatives of the teratogenic drug thalidomide that are approved by the US FDA for the treatment of MM [87,120,177,178,179,180,181,182]. Lenalidomide is the standard-of-care maintenance therapy for patients with standard risk and after ASCT [87,97,183,184,185,186,187,188]. Thalidomide, lenalidomide, and pomalidomide bind to Cereblon (CRBN) [177,189,190,191,192]. CRBN, together with DNA Damage Binding Protein-1 (DDB1), Cullin-4A (Cul4A), and Regulator of Cullins-1 (Roc1), forms a complex called Cullin-RING ligase 4 (CRL4CRBN), an E3 ubiquitin ligase; CRBN functions as a substrate receptor [191]. The drugs inhibit the binding of CRL4CRBN to its endogenous substrate, transcription factor MEIS2, which regulates MM cell survival and sensitivity to anti-MM drugs [191,193]. CRBN in the presence of drugs acquires the ability to target two specific B cell transcription factors, Ikaros (IKZF1) and Aiolos (IKZF3), for proteasomal degradation [194,195,196] (Figure 2 and Figure 3). Ikaros and Aiolos regulate the expression of the key plasma cell differentiation transcription factor, IRF4, which plays a central role in the pathogenesis of MM [197,198] (Figure 3). IRF4 controls B-cell-to-plasma-cell program transition and is essential for Germinal Centers (GC) formation, CSR, and SHM [199]. The AICDA gene encoding AID cytosine deaminase essential for CSR and SHM (Figure 3) is among the IRF4 targets [200]. IRF4 is often overexpressed in MM as a result of mutations, translocations, or other events [201,202,203]. It is of interest that lenalidomide can partially affect the IRF4 level [194,195,197,204]. In addition to that, thalidomide and its derivatives can affect the level of cytokine production since they can inhibit TNF-α and IL-1b production and increase IL-2 and IL-10 production [194,205,206]. Inhibition of NF- κB activation has been reported for IMiDs [207,208]. Lenalidomide can affect the levels of other factors, such as KPNA2 and SALL4 [196,209,210]. It is noteworthy that levels of some of these factors, e.g., KPNA2, can be linked to the overall survival and progression-free survival rates in MM patients [209]. It should be noted that some studies have reported increased risk of SPMs associated with lenalidomide [64,65,74,186,211].
Proteasome inhibitors have the potential to affect multiple aspects of cell homeostasis and metabolism, including DNA damage repair pathways [218,219]. Inhibition of proteasome leads to accumulation of misfolded and unfolded proteins in the endoplasmic reticulum, activates the unfolded protein response, and induces apoptosis; remarkably, the unfolded protein response plays an essential role in antibody-secreting cells [216,220,221,222,223,224,225,226]. A potent reversible PI, bortezomib, has shown more than 60% of positive responses in the initial treatment of MM and ~35% of responses for relapsed MM and is routinely used in schemes for induction therapy and as maintenance therapy for high-risk patients [1,87,88,224,227,228,229,230,231,232,233,234,235,236,237,238]. Two other agents inhibiting proteasome function, carfilzomib and ixazomib, are also used for MM treatment and are superior to bortezomib for the treatment of relapsed or refractory MM [1,88,219,239,240,241,242,243,244]. Such drugs with such a broad spectrum of action have a profound influence on DNA metabolism. Bortezomib affects several DNA repair transactions and can sensitize cancer cells for other DNA-damaging therapeutic drugs. First, bortezomib inhibits the FA/BRCA pathway gene expression and leads to transcriptional repression of FANCD2, BRCA1, BRCA2, and RAD51 genes [245,246]. Second, it reduces monoubiquitination of FANCD2 and polyubiquitylation of γH2AX in cancer cell lines and inhibits the formation of DNA damage foci (BRCA1 and RAD51 foci) in MM cell lines [246,247,248]. Importantly, PIs can affect the NF-κB transcription activation pathway, which is often found dysregulated in MM [207,249,250,251]. Besides that, bortezomib inhibits transcription of the MMSET [252]. MMSET (multiple myeloma SET domain) encodes for a histone methyltransferase involved in transcription regulation and DNA repair. It is also an oncogene that is highly expressed in diverse tumor types. In the case of MM, MMSET is upregulated as a result of the t(4;14) translocation, and its level correlates with the viability of MM cells (Figure 1, see also Section 5.1) [253,254,255]. MMSET, together with NF-kB, is involved in the IRF4 transcription regulation, and bortezomib can downregulate the IRF4 [252]. In addition to the discussed mechanisms, PIs can target telomerase (see Section 6).
Histone deacetylases (HDACs) are involved in many cell processes, from chromatin organization and regulation of transcription to protein stability, trafficking, and degradation. Overexpression of some of HDACs is associated with an adverse prognosis in MM [256]. HDAC inhibitors have been found to inhibit MM cell growth and have demonstrated synergistic activity with bortezomib against MM cell lines [257,258]. Panobinostat is approved for the treatment of relapsed and refractory MM in combination with bortezomib and dexamethasone [259,260]. Panobinostat is a pan-HDAC inhibitor that affects many aspects of cell metabolism [261]. A combination of panobinostat and bortezomib exerts a complementary inhibitory effect on protein degradation systems, including proteasome and aggresome; the latter is inhibited through HDAC6, a master regulator of the cell response to cytotoxic misfolded proteins [256,262]. Panobinostat induces acetylation of histones H3 and H4, activation of caspases 3 and 8, and reduces the levels of the transcription factors IRF4 and MYC [263,264]. MYC is a well-known oncogene that is deregulated in many cancers, including MM. As a part of the B-cell development, program MYC regulates GC formation and is silenced upon switching to the long-lived plasma cell program [198].
Anti-CD38 antibodies have shown efficacy for the treatment of MM in multiple clinical trials in recent years and are included in many treatment schemes [87,88,265,266,267,268]. CD38 is a multifunctional glycoprotein with a high presence on the surface of plasma cells [269,270]. CD38 functions as a cell receptor, adhesion molecule, and ectoenzyme involved in the modulation of many essential cell processes, including immune cell response and signaling [271]. Daratumumab, the anti-CD38 antibody, is approved for MM treatment as a single agent or in combination with other therapies [87,270,272]. Daratumumab induces complement-dependent cytotoxicity and antibody-dependent, cell-mediated cytotoxicity and phagocytosis [269]. Daratumumab significantly reduced the risk of MM progression or death in various groups of patients obtaining different combinations of therapy [273]. The addition of daratumumab to therapy is recommended for high-risk MM patients with refractory/relapsed MM or carrying genetic abnormalities associated with adverse prognosis (see Section 5 for more details) [87]. Remarkably, CD38 is involved in nicotinamide adenine dinucleotide (NAD+) metabolism. While daratumumab seems not to affect NAD+ levels, extracellular and intracellular NAD+ levels may be sensitive to other anti-CD38 molecules such as isatuximab [274,275]. NAD+ is an essential cofactor of many enzymes involved in the regulation of DNA repair and genome stability, including Sirtuins and poly(ADP-ribose) polymerases. NAD+-dependent Sirtuin deacetylase activity is highly sensitive to NAD+ concentrations [271,276]. It is thought that fluctuations in the intracellular level of NAD+ might contribute to genome stability and response to DNA damage, and NAD+-depleting agents might potentiate the benefits of anti-CD38 agents [271,274,277].
Another monoclonal antibody used to treat MM, elotuzumab, targets the signaling lymphocytic activation molecule F7 (SLAMF7), which is highly expressed in plasma cells [1,278,279,280,281]. In addition to that, soluble SLAMF7 has been detected in MM patients but not in healthy individuals [282]. Elotuzumab is not effective in monoregimen against MM but exerts its action in combination with IMiDs [280,281,283]. It promotes an immune response to MM through facilitating natural killer cell-mediated anti-tumor activity [278,284,285]. It is noteworthy that, similar to many other important factors for MM development, the transcription of SLAMF7 is regulated by MMSET [286].
Inhibitors of the antiapoptotic BCL2-family proteins, including BCL2, MCL1, and BCLXL, are in clinical trials, and some of them can be considered for MM treatment. Proteins of the BCL2-family are essential factors of genome stability. BCL2 directly interferes with ribonucleotide reductase activity and affects replication fork progression and DNA damage repair; it also inhibits repair of the DSBs by influencing NHEJ and HR [287]. MCL1 is found at sites of DNA damage and regulates the DNA damage response [288]. Depletion of MCL1 impairs DNA DSB repair and replication re-initiation at stalled forks, thus increasing genomic instability and cell sensitivity to ionizing radiation [289]. MM cell lines are dependent on MCL1 and are sensitive to its inhibition [290,291,292]. Venetoclax is a selective inhibitor of BCL2 approved for treating chronic lymphocytic leukemia (CLL), a BCL2-dependent malignancy of differentiated B-cells [293]. Venetoclax inhibits BCL2 but not MCL1 or BCLXL. However, MM patients with t(11;14) translocation have increased expression of BCL2 and are often sensitive to venetoclax [255,293,294]. It has also been shown that bortezomib can downregulate MCL1 expression and mitigate resistance to venetoclax in some cases [295].
Dexamethasone and prednisone are corticosteroid anti-inflammatory agents routinely used in different combinations to treat MM [87,88]. Dexamethasone blocks NF-kB pathway activation and affects the expression of many genes, including genes implicated in DNA repair pathways and DNA damage response. For instance, it activates the expression of ERCC1 and BRCA2 and downregulates MYC and BCLXL [296,297]. It should be noted that corticosteroids are included in most schemes for MM treatment and may exert a certain effect on genome stability and mutagenesis, especially when used in combination with other drugs influencing DNA metabolism.

2. Predisposition to MM

There is strong evidence for familial clustering of MM and MGUS, which suggests that predisposition to MM is inherited [105,298,299,300,301,302,303,304]. Mutations associated with familial risk of MM development affect the genes that encode for the components of chromatin-remodeling complexes and chromatin-modifying enzymes, DNA repair factors, and RNA processing enzymes as well as cell cycle-regulating factors.
Truncating mutations in the gene LSD1/KDM1A encoding for lysine-specific histone demethylase 1A (LSD1), a component of the NuRD, CoREST, and SIN3A epigenetic regulatory complexes, are associated with predisposition to MM and early onset of the disease [253,305,306,307,308,309]. LSD1 erases mono- and dimethyl groups from H3K4 (active chromatin) and H3K9 (repressed chromatin), and its substrate specificity largely depends on its interacting partners, thus allowing LSD1 to function either as a transcriptional repressor or coactivator of transcription [310,311,312]. LSD1 plays a role in the regulation of stem-cell programs, and it is overexpressed in different cancers, including hematological malignancies [313,314]. In cooperation with CoREST, LSD1 controls hematopoietic cell differentiation [306]. Transcriptional repressor BCL6 recruits LSD1 to the intergenic and intronic enhancers of many genes controlling B-cell development and GC formation as well as to Germline Transcription (GLT) promoters at the IGH locus [315,316]. LSD1 copurifies with MMSET as a part of the SIN3A/HDAC complex, and this interaction might play a role in transcription repression of a number of oncogenes, including telomerase reverse transcriptase TERT and regulator of the p53 tumor suppressor MDM2 [253,309]. Besides that, LSD1 directly participates in the regulation of p53 activity through its demethylation [317]. STAT3 and RB1 are two other non-histone proteins whose activity is regulated by LSD1 [314]. Interestingly, some data suggest that pomalidomide and lenalidomide can activate LSD1, which in turn activates transcription of the cell cycle regulator p21WAF−1 [318].
Pedigree analysis identified that mutations in ARID1A, encoding a key subunit of SWI/SNF chromatin-remodeling complex, are associated with susceptibility to MM [319]. Genes encoding components of SWI/SNF complexes are often found altered in different cancers [320]. Curiously, ARID1A has the highest mutation rate among the genes encoding for the SWI/SNF subunits altered in cancers that can be caused by conditions provoking microsatellite instability [321,322]. There are two main types of SWI/SNF complexes in human cells, classified by their subunit composition: BAF complexes and PBAF complexes; ARID1A is a component of BAF, the complex that interacts directly with topoisomerase IIa (TOP2A) and contributes to the DNA decatenation [323]. Inactivation of BAF functions might result in mitotic defects, chromosomal instability, and polyploidy that are all primary events in cancer development. In addition to that, ARID1A binds to many promoters and enhancers to regulate transcription. For instance, transcription of TP53 is dependent on ARID1A [324]. Note that mutations in ARID1A and mutations in the ARID2, encoding the component of the PBAF complex, occur recurrently as somatic alterations in MM (see Section 3). Waller et al., identified another candidate gene, USP45, as associated with a familial risk of MM [319]. This gene is particularly interesting since it encodes ubiquitin hydrolase that controls the activity of ERCC1, a subunit of the ERCC1–ERCC4 endonuclease (the ERCC1–XPF complex), a core component of nucleotide excision repair that participates in CSR [325,326].
Another gene linked to MM development, DIS3, is also one of the top somatically mutated genes in MM (see Section 3) [28,53,82,86,327,328,329]. The presumptive loss-of-function mutations in DIS3 have been reported in about 2.6% of familiar cases with MGUS and MM [330]. DIS3 encodes for the catalytic 3′-5′ exonuclease subunit of the exosome complex, which is implicated in processing RNA transcripts and regulating mutagenesis and recombination at the IGH locus [331,332]. It is tempting to speculate that DIS3 mutations might be implicated in aberrant CSR, leading to primary genomic rearrangements at the IGH locus (see Section 5.2 for more discussion).
WES analysis of the family with increased prevalence of MGUS and MM allowed identifying a missense mutation in the EP300, encoding transcriptional coactivator p300, as a likely causative variant [333]. Recurrent somatic mutations in EP300 in MM were also described (see Section 3). In addition to that, rare germline variants associated with MM familiar cases were reported for CDKN2A [334,335]. CDKN2A encodes for several transcripts: p16-INK4A and p14ARF encoding for proteins acting as tumor suppressors for different cancers. The p16-INK4a belongs to the INK4-family of specific inhibitors of CDK4/6 kinases, controls G1-to-S progression, and is an essential component of cellular senescence program [336,337]. Another transcript, p14ARF, is a result of alternative splicing, and the synthesized protein plays a role in TP53 activation by binding to MDM2 [338,339]. Methylation of the regions corresponding to p14ARF and p16-INK4A could represent early events in the pathogenesis and development of plasma cell disorders and were reported in MM [340,341]. Expression of p16-INK4A is often absent in MM [342].
It should be noted that LSD1, ARID1A, USP45, DIS3, and EP300, in addition to the mechanisms discussed above, can be implicated in the telomere maintenance pathway, and thus might influence the predisposition to MM through telomere structure and length regulation (see discussion in Section 6).
More than 100 different SNPs were also reported in GWAS studies to associate with the risk of MM development [343,344,345,346,347,348,349,350,351] or MGUS [352,353]. Genes and pathways affected by these SNPs are proposed to be susceptibility candidates based on linkage disequilibrium distance and/or altered gene expression. In many cases, they affect non-coding regions or lie thousands of base pairs away from the candidate gene. In total, genes affecting several biological pathways have been tightened to MM development risk, including histone modification and chromatin remodeling, transcription, and co-transcriptional RNA maturation, IRF4-MYC regulatory network, B-cell differentiation, genome stability, and telomere maintenance (see Table S3). A recent approach taking advantage of a transcriptome-wide association study allowed for the expansion of GWAS analysis and identified new MM risk genes, including DNA/RNA-editing cytosine deaminases APOBEC3C, APOBEC3D, APOBEC3F, APOBEC3G, and APOBEC3H at 22q13.1, responsible for immunity, and RNF40 at 16p11.2, involved in DSB-repair [354].
It should be noted that some SNPs identified in GWAS studies were linked to chromatin remodeling and epigenetic regulation that, in connection with the risk variants observed for family cases, highlights the importance of chromatin organization and regulation for MM development. One example is SNPs at 7q36.1, presumably affecting the expression of SMARCD3, encoding a component of the BAF complex [344,346,347,355]. Several other nucleotide variants affect components of the PRC1 and PRC2, the two major Polycomb group (PcG) repressive complexes controlling cell-specific transcriptional programs via histone modifications [356,357]. Thus, SNPs at 22q13.1 localize to the CBX7 locus encoding a component of PRC1, the complex that facilitates H2AK119 ubiquitination and plays a role in lymphogenesis and hematopoietic malignancies [345,358,359,360]. H2AK119 ubiquitination at sites of DNA breaks has been linked to DSB repair [357,361]. Overexpression of BMI1, encoding for another subunit of the PRC1 complex that stimulates the ubiquitinase activity of PRC1 toward H2AK119 and promotes DSB repair, has been observed in myeloid malignancies [361,362]. BMI1 is essential for MM cell growth, and its depletion sensitizes MM cells to bortezomib [363,364]. It worth noting that another SNP, rs34229995, associated with MM localizes to the regulatory region of JARID2 encoding the subunit of the PRC2 complex [344,365]. JARID2 binds to H2AK119Ub and facilitates trimethylation of H3K27 by the PRC2, thus promoting transcriptional repression [357]. Curiously, PcG complexes are known negative regulators of CDKN2A expression, the gene that has been linked to familiar MM cases [366,367]. The involvement of CDKN2A in MM development is additionally confirmed by the association of rs2811710 located in the regulatory region of the CDKN2A with MM risk [344]. Several SNPs at 7p15.3 are linked to the increased expression of CDCA7L [352,368,369,370]. CDCA7L affects the genome-wide methylation level, and it is a target gene of MYC; it is also involved in CCND1 upregulation in glioma [371,372].
Another group of SNPs associated with MM development is linked with transcription regulation and RNA processing. Thus, rs4325816 is located in the intron of the SP3 gene encoding transcription regulator, which is among the bortezomib targets [373]. SNPs located at 19p13.11 are located near the KLF2 gene, which is listed among hotspots for structural variations in MM [83]. KLF2 is a transcription factor that plays a role in maintaining B-cell quiescence and regulating IRF4, BLIMP1, and AID levels; its expression is induced after pre-BCR signaling and is maintained until B-cell activation [374]. Bortezomib upregulates KLF2 [375]. SNPs at 5q15 are linked with ELL2 encoding for the elongation factor of RNA polymerase II, which regulates immunoglobulin mRNA processing in plasma cells [376,377,378]. ELL2 travels with the RNA polymerase II across the IGH μ- and γ-gene segments and mediates the association of polyadenylation-cleavage factor CstF-64 with the RNA polymerase II and switching to a secretory-specific poly(A) site [378].
Regulation of DNA repair and B-cell-specific genome rearrangements, such as CSR and V(D)J recombination, seem to associate with MM development. Polymorphisms in the AICDA gene were linked with better outcomes in MM patients [379]. Decreased MM risk has also been reported for SNPs located near LIG4, encoding the DNA ligase IV essential for V(D)J recombination [379,380]. On the other hand, SNPs located in the regulatory regions of XRCC5 encoding Ku80 and XRCC4 encoding a factor associating with DNA ligase IV conferred susceptibility to MM [381]. A missense mutation in the RFWD3 encoding a RING-type E3 ligase that belongs to a FA group and is essential for HR repair, replication fork restart, and translesion DNA synthesis predisposes to MM [344,382,383,384,385,386]. Several risk SNPs associate with TNFRSF13B and TOM1, the genes implicated in immunoglobulin class-switching regulation [343,344,368,377,387,388]. Autophagy and intracellular trafficking play an important role in plasma cell differentiation and immunoglobulin production; consistently, the defects in autophagy have been linked to MM development [216,344,389]. It should be noted that autophagy-deficient Atg5−/− plasma cells have higher expression of BLIMP1 and IGH genes [389,390]. Several SNPs affect genes with the proposed centrosomal role, including the ULK4 and CEP120, and might potentially impact chromosome stability [348,368,391,392,393]. In addition to the discussed loci, multiple SNPs affecting genes associated with the telomere maintenance pathway have been linked to MM predisposition risk (see Section 6 for discussion).
Altogether, the set of risk mutations and SNPs associated with MM predisposition highlight the role of chromatin structure and epigenetic regulation, DNA repair, transcription, and RNA processing, as well as B-cell-specific processes necessary for antibody synthesis and maturation as significant factors influencing disease development. It is plausible that B-cell-specific IGH locus reorganization might predispose to certain types of chromosomal rearrangements and aberrations that initiate MM development. It should be noted that V(D)J and CSR are mutagenic processes per se and might initiate a cascade of genomic alterations, ultimately leading to carcinogenesis (see Section 5.2 for more discussion).

3. Mutational Landscape of MM and Its Precursors

Complexity and heterogeneity are hallmarks of the MM cells’ genomic landscape. The distribution of mutation loads is highly variable in both MM and other hematological tumors. However, MM (as well as other B-cell neoplasms) tend to possess more point mutations than other hematological malignancies, though less than cancers associated with higher mutagenic exposures, for example, melanoma, which has roughly 10 times more mutations compared to MM [394]. The first NGS studies of MM focused on detecting mutational profiles. The scope of the following studies was expanded to identify specific genetic alterations for different stages of the disease, find driver mutations, deduce MM-specific mutational signatures, and locate pathways involved in the progression from early stages toward MM (Table S1). The tumor-normal pair approach compares the genomes of tumor and germline variants and allows for finding mutations present in MM only.
A total of about 60 significantly mutated genes (i.e., the genes that are mutated more often than expected by chance, as determined by a specific algorithm) have been identified in MM cells [28,61,81,82,86,328,395,396]. Most mutations affect the following pathways: signaling RAS/MAPK, NF-kB, MYC, DNA repair, RNA interaction, chromatin regulation and epigenetic mechanisms, transcription, cell cycle regulation, plasma cell differentiation, cellular adhesion, and motility. Genes recurrently mutated in MM and the corresponding pathways implicated in MM development are summarized in Table S2.
Mutations affecting the RAS/MAPK pathway prevail in newly diagnosed patients and were found in 19–50% of patients [28,81,82,86,328,397,398]. Mostly, KRAS, NRAS, and, to a lesser degree, BRAF mutate in different clones. KRAS and NRAS mutations are nearly always mutually exclusive but can coincide with BRAF mutations [28,328]. These events can be both clonal and subclonal, meaning they can arise early and late in clone evolution, correspondingly [28,328]. Mutations in other genes, such as NF1, PTPN11, and RASA2, leading to activation of the RAS/MAPK pathway, were also reported [86,328,398] (Table S2). The NF-κB pathway participates in apoptosis, differentiation and proliferation of cells, immunity, inflammation, and lymphopoiesis, is affected in 40% of MM cell lines and 12–20% of MM patients, and is represented by mutations mostly in TRAF2, TRAF3, CYLD, NFKB2, NFKBIA, and LTB genes [28,82,86,328,399,400]. Despite their high prevalence in MM and apparent association with tumor progression (i.e., driver mutations), mutations in RAS/MAPK and NF-kB pathways in many cases showed no prognostic impact on survival [82].
Mutations in some of the DNA repair pathway genes are also frequently found in MM patients. In most cases, they affect genes encoding factors facilitating DNA damage sensing: the p53 tumor suppressor and the ATM/ATR kinases facilitating p53 phosphorylation [82,86,328,401]. Frequency of detected TP53 changes correlated with the disease stage: On average, 5–8% of MM patients carry TP53 point mutations at the diagnosis, but up to 25% at the advanced stages of the disease [402]. HUWE1 encoding E3 ubiquitin ligase involved in DNA damage response and DNA repair is mutated in a significant cohort of MM patients, contributing to tumor growth and MM treatment [86,403]. Mutation in other DNA repair pathways, e.g., HR, FA/BRCA, NER, BER, and MMR, occur in MM patients more rarely, with a frequency of about 1% or less [82,86,404,405,406]. As was discussed in Section 1.2, mutations in some of these genes or deregulation of their expression may associate with resistance to chemotherapeutic drugs, e.g., melphalan [160,404].
Mechanisms of RNA processing have a significant impact on MM development. The DIS3 gene is found recurrently mutated in MM patients with a frequency of about 8–11% at the time of diagnosis [28,53,81,82,86,327,328,329]. DIS3 mutations are also associated with inherited MM risk (see Section 2). Somatic DIS3 mutations in MM are mostly clonal; however, their subclonal accumulation has also been noted and has been associated with shorter overall survival [28,327,328]. Another significantly mutated gene in MM is FAM46C, which is affected in 5–13% of cases and is deleted in 20% of patients [29,81,82,86,407,408]. FAM46C encodes for a non-canonical, poly(A)-polymerase controlling mRNA stability and facilitating B-lineage-specific onco-suppressor functions [409,410]. FAM46C is upregulated during plasma cell development and contributes to immunoglobulin heavy and light chains’ mRNA poly(A) length, stability, and immunoglobulin production [409,411,412]. Mutations in FAM46C are specific to MM and are infrequent in other cancers. They are associated with worse overall survival in MM patients and are predominantly found in hyperdiploid tumors [29,328,407]. A minor proportion of MM cases harbor mutations in the other genes encoding for RNA interacting proteins [29,398,406] (Table S2). Mutations in the RPS3A are common for hematological malignancies [413]. In addition, mutations in the genes encoding for ribosomal proteins RPL5 and RPL10 have been noted and RPL5 is frequently deleted in MM [414].
Mutations affecting chromatin regulation and epigenetic mechanisms are common for MM. One interesting group is represented by genes encoding linker histones. Missense mutations clustered in the regions corresponding to the highly conserved globular domain of the linker histones encoded by genes HIST1H1B, HIST1H1C, HIST1H1D, HIST1H1E, HIST1H2BK, and HIST1H4H were found in MM cells [61,82,86,398,415]. Linker histones facilitate nucleosome–nucleosome contacts and contribute to the formation of nucleosome clutches representing higher-order chromatin interactions. They also play a role in establishing Topologically Associated Domains (TADs) and modulation of chromatin topology within the nucleus [416]. Affecting linker histones’ function can have an impact on chromatin structure, genome stability, and gene expression, which ultimately may influence the rate of genetic and epigenetic changes in MM.
Another group is represented by genes encoding histone modifiers. Genes of histone lysine methyltransferases and demethylases frequently alter in MM, and their mutations may account for epigenetic deregulation associated with disease progression and resistance to therapy [417]. Most frequently, mutations are observed in the genes of the MLL family, MLL (KMT2A), MLL2 (KMT2B), MLL3 (KMT2C), MLL4 (KMT2D), and MLL5 (KMT2E), involved in methylation of the H3K4 residue, a mark of transcription activation [81,86,415]. Other mutated methyltransferase genes include, but are not limited to, EHMT2, SETDB1, and SETD2, which contribute to methylation of H3K9, H3K27, H3K36, and MMSET, which contribute to methylation of H3K27, H3K36, and H4K20 [29,86,415]. MMSET is a master regulator of many pro-oncogenic pathways, and its upregulation or gain-of-function mutations have important prognostic significance (see Section 1.2) [418,419]. Mutations affecting demethylases are most frequently found in UTX (KDM6A), KDM3B, and KDM5C, which encode enzymes removing methyl residues from methylated H3K27, H3K9, and H3K4, correspondingly. These methylated sites are considered marks of silent or inactive chromatin [29,86,328,329,415,420]. Moreover, MM genomes harbor mutations in the genes EP300 and CREBBP(CBP) [86,415]. EP300 and CREBBP are closely related paralogs encoding major lysine acetyltransferases in metazoans, functioning as transcriptional coactivators and tumor suppressors [421]. The p300 contributes to the acetylation of the H3K27 and H3K18 histone residues, the modifications that associate with active promoters and enhancers. In addition to that, p300 acetylates many non-histone targets. Mutations in EP300 are associated with many different cancers and are often found in patients with myelodysplastic syndromes [421,422]. Chemicals that bind to bromodomains of p300 and CBP abrogate the viability of MM cell lines by interfering with the IRF4 transcriptional program [423]. Mutations in NCOR1, encoding a co-repressor recruiting and activating HDAC complexes, were also found [53,86].
Chromatin remodelers participate in the regulation of chromatin organization essential for DNA replication, DNA repair, transcription, and other processes and play an important role in genome stability. It is not surprising that MM genomes also possess mutations in the genes encoding chromatin remodelers and their interacting partners, including CHD2, CHD4, ARID1A, ARID2, ATRX, and ZFHX4 [9,82,86,415]. Noteworthy, ARID2 is a neosubstrate of CRL4CRBN induced by IMiDs, and it is effectively targeted for degradation by pomalidomide [424]. Mutations in the ARID1A also associate with the risk of MM development (see Section 2), highlighting the role of chromatin regulation in shaping the MM genome.
A special niche is taken by genes encoding DNA modifiers. IDH1, IDH2, DNMT3A encoding DNA methyltransferases and TET2 (Ten-Eleven Translocation 2) encoding for methylcytosine dioxygenase are among frequently mutated genes in MM [60,86,415]. The DNA modification process seems to play an essential role in transitioning from MGUS to MM. A global transition to CpG islands’ hypomethylation in genic regions has been noted after a shift from a non-malignant state to MM. The event affected the expression of many genes and was linked to the downregulation of the DNMT3A [425,426,427]. The hypermethylated regions in MM cells mainly concentrate in B-cell-specific regulatory regions (see Section 5) [428]. A correlation between different structural rearrangements in MM and methylation patterns has also been described [425,429]. Interestingly, a global level of a 5-hydroxymethylcytosine mark is also decreased in MM compared to normal plasma cells [430]. TET2, which facilitates oxidation of methylcytosine in the DNA, thus contributing to the production of 5-hydroxymethylcytosine, is frequently deregulated in myeloid malignancies, and TET2 inactivation has been proposed to play an initiating role in age-related hematological cancers [422,431,432,433]. Importantly, TET enzymes bolster the expression of AICDA, thus augmenting CSR [434]. Both DNMT3A and TET2 mutations are enriched during the relapse stage of MM, which correlates with chemo-resistance [415].
Among genes encoding transcription factors, frequent mutations in MM have been observed in IRF4, BLIMP1(PRDM1), and XBP1 [28,61,81,82,86,328,398]. As discussed above, BLIMP1, XBP1, and IRF4 are transcription factors expressed in antibody-secreting plasma cells and are important for plasma cell commitment (see Figure 3). Mutations in the cis-regulatory elements of other important players of B-cell development, e.g., PAX5, BCL6, and PAXIP1 (PTIP), were also described in MM and may affect their expression [60,398]. Other transcription factor genes whose mutation are found in MM are EGR1, MAX, ZNF208, MAF, MAFB, IKZF1, and IKZF3 [28,29,61,81,82,86,328,398]. EGR1 mutations are associated with hyperdiploidy and often occur in WRCY DNA sequence motifs, implicating SHM and AID as causative factors [28,82,435]. Mutations in MAX, IRF4, and EGR1 may affect the expression level of the MYC oncogene. Despite these facts, IRF4 and EGR1 mutations positively impact the survival of MM patients [82].
Several cell cycle pathway genes mutate recurrently in MM. The most essential are CCND1, CCND3, RB1, CDKN2C, and CDKN1B [28,60,61,82,86,161,395]. Mutations in CCND1 are an early event in the development of MM and associate with t(11;14) translocation (see Section 5.1). CDKNK2C and RB1 are potent cell-cycle regulators, and their inactivation correlates with a high proliferation index.
Dynamic changes and mutational processes involved in MM pathogenesis are revealed by analysis of genomes at different stages of the disease, starting at diagnosis, progression, and relapses or post-treatment time points [29,61,63]. The genetic determinants of the MM evolution from MGUS and SMM stages (Figure 1) are much less understood. Several WES and WGS studies characterized the spectrum of mutations and structural rearrangements of MGUS and SMM to find the genetic patterns of progression from premalignant stage to MM [20,48,51,62,436,437,438,439]. The genomic dataset of both MGUS and SMM contains fewer mutations than MM, but mutations of KRAS, NRAS, DIS3, HIST1H1E, EGR1, LTB, and CCND1 occur both in MGUS and MM [20,51,52,62]. The mutational landscape of SMM is similar to newly diagnosed MM, although it varies in groups with different progression risks to MM [48,62]. Mutations of NRAS, KRAS, BRAF, DIS3, FAM46C, TRAF3, TP53, ATM, LTB, EGR1, RB1, MAX, CDKN2A, and other genes were found in SMM patients [47,48,51,56]. Importantly, no mutations in the DNA repair genes such as TP53, ATM, and ATR were identified in stable MGUS patients, and biallelic inactivation of TP53, RB1, DIS3, MAX, and CDKN2A were rare events for low-risk SMM, suggesting that these abnormalities are associated with tumor progression [48,51,52,62].
Several studies have revealed genomic changes in response to drug therapy [25,53,54,60,440,441,442,443,444]. As was discussed in Section 1.2, the CRBN protein is the main target of IMiDs, such as thalidomide, pomalidomide, and lenalidomide, used in MM therapy. Mutations in the CRBN and in the genes acting in this pathway, such as IKZF1, CUL4A, and IRF4, were more frequently found in patients refractory to IMiDs and associated with worse outcomes [329,441,443]. A shorter duration of therapy with IMiDs was proposed to result in fewer mutations in cereblon pathway genes [209,443]. PIs are widely used in many MM treatment regimens, but almost all patients develop resistance to these drugs over time. It has been shown that treatment with PIs may cause accumulation of mutations in the genes encoding components of proteasome such as PSMB5, PSMB8, PSMD1, and PSMG2 [442]. Recent studies suggest that MM at relapse has a more complex genetic landscape compared to primary tumors and highlights the biological role of TP53 inactivation and certain structural genome rearrangements (e.g., 1q gain, del(17p), and MYC translocations, see Section 5.1) rather than mutations in genes associated with resistance to PIs or IMiDs in acquired resistance to therapy [25,54,60]. Notably, the response to IMiDs could also be influenced by the naturally occurring polymorphism in the CRBN gene region and by the epigenetic modifications of the CRBN regulatory elements [445,446].

4. Mutational Signatures in MM Genomes

Tumor cells accumulate multiple mutations during cancer initiation and progression. Types of mutations, DNA sequence context where they happen, and the distribution of mutations in the genome help deduce what molecular processes operated during tumor evolution [447,448]. It is clear that the mutational landscape of a tumor is shaped by the concerted action of many factors, including therapeutic drugs: There are signatures associated with age, smoking, UV light, various DNA repair defects, APOBEC deaminases, chemotherapy, and others [448]. At least five signatures are seen in MM (Figure 4). High-dose chemotherapy contributes to the mutational landscape of MM, increasing the mutational burden and producing new mutational signatures [54,59,63,145,146]. Thus, melphalan signature has been found in the transcribed strand implementing transcription-coupled NER on damaged DNA [63,146]. Three mutational signatures common in cancer genomes were attributed to the action of the AID/APOBEC family of cytosine deaminases [449]. AID/APOBEC is a superfamily of enzymes that convert cytosine to uracil and consists of two subfamilies, AID and APOBEC3 [450]. AID (activation-induced cytosine deaminase) edits DNA in immunoglobulin genes in the process of SHM and CSR in the maturating B-cells. Proteins of the APOBEC3 subfamily mediate the restriction of viruses by deaminating viral cDNA. AID/APOBEC proteins are active on ssDNA and can act processively, thus generating clustered mutations [451]. Individual APOBEC protein family members possess intrinsic deamination sequence specificity, thus allowing for attributing the mutations found in cancers to the action of specific APOBEC based on the sequence context of the changes. It has been proposed that AID/APOBEC processivity and availability of long ssDNA regions in tumor evolution give rise to a phenomenon called kataegis–localized hypermutated regions [452,453,454].
Signatures of AID/APOBEC deaminases-related mutagenesis are found in many cancers, including MM [63,79,82,394,395,435,447,448,456,457,458]. The first paper on the classification of mutational signatures in cancer [394] reported two signatures in multiple myeloma: signature 2, which is attributed to the activity of APOBEC3 proteins, and signature 5, which has an obscure mechanism, possibly related to smoking. Further in this review, mutation signatures are referred to as SBS (for Single Base Substitutions), according to the newest classification [448] (see https://cancer.sanger.ac.uk/signatures/sbs/, accessed on 20 October 2021). One of the early papers also reported detecting clustered mutations, kataegis, in the APOBEC3-specific sequence context in MM [459]. AID preferentially deaminates cytosines in WRCY sequence context (mutated base underlined), where W = A/T, R = A/G, and Y = C/T. The resulting uracil in DNA (WRUY) can be replicated through emergence of C->T transitions and WRTY (AID SHM signature, SBS84) [448] or by inducing a series of DNA repair events involving error-prone DNA polymerases η and θ, leading to A->C/G substitutions in WA motifs (SBS85, Figure 4) [435,460,461,462]. In the hematological malignancies’ literature, SBS85 is often referred to as a non-canonical AID signature [63,398].
Proteins of the APOBEC3 subfamily have a strong preference for TC sequences, thus causing TC->TT and TC->TG mutations (the latter likely due to the activity of REV1 translesion DNA polymerase) [463], giving rise to signatures SBS2 and SBS13 (Figure 4). These signatures are often referred to as “APOBEC” signatures, but they are different from the above AID signature, a member of AID/APOBEC family enzyme (but different subfamilies). With the use of a slightly longer sequence context, it is possible to separate SBS2 and SBS13 signatures into APOBEC3A- and APOBEC3B-specific signatures [448,464,465].
One of the first reports on whole exome/genome sequencing (WES and WGS) of MM revealed a higher mutation frequency in noncoding and intronic regions, as compared to protein-coding regions, which was at least partially explained by the action of transcription-coupled repair [81]. In addition, 5’-UTRs and the beginning of the coding sequences of genes known as SHM targets (immunoglobulin genes) and off-targets (genes that do not normally undergo SHM, but tend to be mutated in cancers due to abnormal AID recruitment, for example, BCL6 and MYC) were also enriched with mutations [81,466]. Mutations at the beginning of the genes (encompassing both promoter regions and beginning of CDSs) were found in the typical AID consensus motif WRCY, implicating that AID was a mutator in studied MM cases [28].
The following WES and WGS studies of MM elaborated on the original findings. Using WES, signatures attributed to three mutational processes have been detected in MM: (1) APOBEC3-specific signature, sometimes with mutation clustering, kataegis; (2) SHM-like mutations (in C-T pairs and A-T pairs), often clustered; and (3) C->T mutations in CpG motifs, age-associated signature found in many cancers that are caused by deamination of methylated cytosine, which is generally considered to be spontaneous, but can also be AID-mediated [29,456,467]. Clusters of mutations were often found close to translocation breakpoints, with AID-type mutations generally found in translocations involving the IGH locus, the natural target of SHM, while other translocations sometimes possessed APOBEC3-like mutations [456]. More recent reports involving WGS allowed for the characterization of MM genomes with improved resolution. Signatures of APOBEC3′s, AID, and SHM-Polη (“non-canonical AID”, SBS85) have been found, with the IGH locus possessing kataegistic regions containing all three mutational signatures [47]. Similar results were obtained in a study combining WES and WGS, and more AID-type mutations have been found in the driver genes, indicative of the undergoing selection [395]. The combined results indicate that SHM-like, AID-initiated mutational events usually happen in the early stages of myeloma evolution when B-cells undergo antigen-dependent maturation in GC [63,468,469]. Later in myeloma evolution, APOBEC3-induced mutagenesis predominates. Interestingly, myelomas with specific translocations (t(14;16)/MAF, t(14;20)/MAFB, and t(8;14)/MAFA the “maf” translocation group, see Section 5.1) possessed an overall higher number of mutations and a larger proportional contribution of APOBEC3 signature compared to tumors with other translocations [63,395,398,456]. It turns out that the overproduction of transcription factors MAF, MAFB, and MAFA as a result of translocations in these tumors leads to the increased expression of APOBEC3s, which in turn causes increased APOBEC3-dependent mutagenesis [456]. Moreover, by separating APOBEC3A and APOBEC3B mutational signatures [464], Rustad et al. [63] concluded that APOBEC3A, but not APOBEC3B, is the main source of mutations in these highly mutated “maf” translocation type myelomas. Altered relative proportions of these mutational signatures in tumors with higher APOBEC3A-to-APOBEC3B ratio reinforced these findings [62]. Overall, the increased APOBEC3 activity is more frequent in the progressive disease. However, stable MM precursor clones may possess “maf” translocations and have an increased APOBEC3A:3B ratio [62,470]. Clearly, more research is needed to decipher the relation between “maf” translocations, APOBEC3A, APOBEC3B, and MM development. Further studies using more sophisticated algorithms of separating highly similar mutational signatures [465] might further clarify these complex processes.
Taken together, MM genomes’ sequencing data combined with analysis of specific mutational signatures favor a model where AID-dependent SHM machinery participates early in MM evolution, inducing mutations in driver genes and leading to a bias towards noncoding regions and the beginnings of the genes. Later, APOBEC3-dependent mutagenesis is continuing to shape tumor genomes, with the age-related CpG deamination process stably contributing during the whole process of tumorigenesis [63]. Finally, APOBEC-induced mutagenesis is associated with MM progression from its precursors [48,51,62]. It has been proposed that testing for APOBEC activity at the time of diagnosis might suggest better, more personalized treatment options [438]. Specifically, since MM patients with high APOBEC activity have a poor prognosis, it was suggested they might benefit from more aggressive treatments. It is worth noting that approaches are being developed to target APOBEC enzymes [471].

5. Structural Variations in the MM Genomes

5.1. Driving Genomic Rearrangements in MM

MM is characterized by karyotypic changes in the malignant plasma cells. Fluorescent in situ hybridization along with classic cytogenetic analysis has been used for years to analyze the karyotype in MM [472,473,474,475]. Recent advances in NGS sequencing of tumor genomes allowed more precise and accurate characterization of the genomic landscape in MM, determination of complex rearrangements, and tracking the chronology of events [79]. Various types of karyotype instability observed in MM include chromosomal gains or losses, translocations, and complex genomic rearrangements [78,476]. Karyotype instability is currently viewed as an early event in the development of MM, driving the development of the disease, and represents an important prognostic factor. Two major classes of primary karyotypic changes in MM are hyperdiploidy and IGH-translocations.
Trisomies of several odd chromosomes, most often chromosomes 3, 5, 7, 9, 11, 15, 19, and 21, are found in tumor cells in approximately 60% of patients with MM, and this is referred to as a hyperdiploid state [61,477,478]. The type and number of trisomies have prognostic significance [479]. Recent data suggest that the acquisition of the hyperdiploid state is a stepwise process, which begins at the early stages of the disease development [61,79,478]. Plasma cells in most patients with MGUS carry numerical abnormalities of at least one of chromosomes (most often 6, 9, 13, 15, 17, 19), and up to ~50% are classified as hyperdiploid [57,478,480,481]. More chromosomes are gained as the disease progresses, and some patients acquire whole-genome duplications at the extreme end of the disease evolution [61,478]. The most common chromosomal translocations in MM involve chromosome 14 and specifically occur at the IGH locus at 14q32.33. This locus encodes for heavy immunoglobulin chains and undergoes multiple changes during B cell differentiation, apparently providing an important source of genomic instability in MM. The chromosomal aberrations at the 14q23.33 are regarded as early stages of MM development as they are found in nearly 50% of patients with MGUS and SMM [57,482]. Five major types of chromosomal translocations involving IGH in MM are t(4;14), t(6;14), t(11;14), t(14;16), and t(14;20) [483,484]. In all these cases, a specific oncogene on the partner chromosome is placed under the IGH enhancer control, resulting in an upregulation of its expression. Translocations t(11;14) and t(4;14) are the most frequent and usually are found in 15–20% of patients with MM each, while t(14;16), t(6;14), and t(14;20) are less common and each occur with a prevalence of 5% or less.
Translocation t(4;14) involving the 4p16 region is associated with adverse prognosis and results in upregulation of FGFR3 and MMSET genes in most carriers [78,483,485,486]. This translocation is hard to identify by G-banding because the breakpoints on chromosome 4 occur in the telomeric region close to the FGFR3 that becomes associated with the 3′ enhancer(s) on derivative chromosome 14 [484]. On chromosome 14, breakpoints usually map to the switch regions that play an important role in immunoglobulin isotype switching (see Section 5.2), indicating that the mechanism of this translocation is connected to CSR [484]. FGFR3 overexpression results in the activation of the transcription factor STAT3 and consequent upregulation of its target BCLXL gene, which is functioning in suppressing apoptosis and promoting MM development [487]. Besides BCLXL, STAT3 regulates the expression of other anti-apoptotic genes, MCL1 and BCL2, key oncogenes including MYC and CCND1, and telomerase reverse transcriptase TERT [488,489]. Interestingly, although mutations in STAT3 have not been reported in MM, STAT3 has been found to be constitutively activated in plasma cells isolated from MM patients as well as in MM cell lines [490,491,492]. In addition to that, IL6-STAT3 signaling is a key regulator of the B-cell differentiation program, and STAT3 is specifically required for the transition to antibody-secreting program [214,493,494]. FGRF3 is also frequently found mutated in primary MM tumors and MM cell lines carrying t(4;14), which also may lead to STAT3 and MAPK activation [495]
Upregulation of MMSET leads to an increase of active chromatin mark H3K36me2, decrease in the repressive mark H3K27me3, and global increase in chromatin accessibility [418,419,496]. MMSET mediates p53 degradation and it is involved in the regulation of expression of DNA repair genes; it enhances DNA damage repair, cell proliferation, and resistance to melphalan [497,498]. MMSET regulates DNA damage-induced histone H4K20 methylation, 53BP1 foci formation, and DSB repair [499]. It also interacts with PCNA, and it is required for replication licensing and normal S-phase progression [500]. MMSET promotes NHEJ at unprotected telomeres and, therefore, might contribute to the generation of complex genomic rearrangements [501]. MMSET also contributes to AID-mediated DSBs at switch regions at the IGH locus [502].
Translocation t(11;14) attaches the cyclin D1 gene CCND1 to the IGH enhancer. On 11q13, the breakpoints are dispersed over 330-kb intervals centromeric to the CCND1 gene [484].
Translocation t(6;14) involving 6p21 juxtaposes cyclin D3 CCND3 gene to the IGH enhancer. The breakpoints are within 150-kb intervals centromeric to the CCND3 [484].
Several primary IGH translocations form the so-called “maf” group because they result in upregulation of the MAF transcription factors. Translocation t(14;16) juxtaposes the IGH locus to the 16q23 with breakpoints located in the last exon of the tumor suppressor WWOX and affects the neighboring oncogene MAF expression. Importantly, WWOX is located in the same band 16q23.2 with a highly unstable region, a common fragile site (CFS) FRA16D [503]. FRA16D CSF is known by association with genomic instability in different types of cancer, and WWOX downregulation is thought to play a role in tumor development [503,504]. A translocation involving 20q12, t(14;20), results in the upregulation of the MAFB gene, a paralog of MAF [505]. The third translocation belonging to the “maf’ group, t(8;14), juxtaposes the IGH locus to another paralog of MAF, the MAFA gene, at 8q24 and it is rarely found in newly diagnosed MM patients [506]. These translocations are characterized by high APOBEC activity since MAF factors regulate APOBEC expression [63,456] (see Section 4). A few other translocations involving the IGH locus in MM involve 1q21, 6p25, 20q11, 21q22, and 22q12 and are infrequent in primary patient samples; they are mostly found in cell cultures [484,506,507].
Significant association of translocations with increased mutagenesis in certain genes has been observed. For instance, t(11;14) is associated with mutations in CCND1 and IRF4, t(4;14) group is associated with mutations in FGFR3, PRKD2, and DIS3, t(14;16) coincides with mutations in MAF, ATM, BRAF, TRAF2, EP300, and DIS3, and t(14;20) accumulates mutations in MAFB [29,61,81,82,86,395,456]. The genes juxtaposed to the IGH regulatory regions because of rearrangement may harbor cluster mutations as a result of SHM driven by AID or other factors spreading to new genomic sites. It is also possible that new genomic locations and increased expression of the affected genes may contribute to the mutation rate. Note that some of the listed genes are not affected directly by the translocations and other mechanisms can account for accumulation of mutations in these regions. Interestingly, the role of IGH enhancers in plasma cell development and oncogenesis has led to an idea about the possibility of regulating their activity through small molecules [508]. One of the selected molecules in this research inhibited the expression of the translocation-induced oncogenes and showed efficacy in trials with MM cell lines.
Other events that can be tracked as early as the MGUS state include a gain of 1q, 13q deletion, and, in rare cases, 17p deletion [20,57,61,478,509]. Some karyotypic abnormalities detected in MM, hyperdiploidy, or a few types of IGH translocations have a neutral or neutral-to-poor prognostic significance if present in combination with other genomic changes. Well-established, high-risk prognostic factors in MM include the translocations t(4;14), t(14;16), or t(14;20), the deletion of 17p, and the gain of the 1q arm [87,510,511,512,513,514,515,516].
Chromosomal aberrations involving chromosome 1 are among the most common cytogenetic alterations in MM and include deletions of 1p and amplification of 1q. Several key genes connected to MM etiology are located in the regions affected by some of these aberrations and are often subject to biallelic loss. Thus, tumors in ~4% of MM patients are characterized by a biallelic loss of FAM46C resulting from homozygous 1p12 deletions, point mutational inactivation of the FAM46C, or both [86,407]. The biallelic loss of FAM46C was linked to a shorter overall survival through activation of the PI3K-Akt signaling pathway, which augments MM cell growth and survival [407,517,518]. The CDKN2C gene located at 1p32.3 is found deleted in 9–33% of MM tumors that are associated with worse overall survival in MM patients who received ASCT [86,161,407,476,519]. Biallelic inactivation of CDKNK2C is frequently observed in MM and is associated with disease progression and has a poor prognosis [407,476].
A heterogeneous class of 1q gain/amplification events is detected in about 40% of patients at the time of diagnosis, and it is one of the most frequent events in MM associated with poor prognosis [520,521,522]. Gain of the 1q may be a result of trisomy or a whole-arm/partial translocations of 1q to the same chromosome or another chromosome. Translocations of the 1q arm may involve several recipient chromosomes, for instance, chromosomes 5, 8, 12, 14, 15, 16, 17, 19, 21, and 22, and most often occur within their telomeric and pericentromeric regions [523,524]. Translocations involving several recipient chromosomes are called “jumping”. The breakpoints on donor chromosome in the case of MM are often localized to the pericentromeric region of the long arm of chromosome 1, with a remarkably high frequency in the regions 1q10–1q11 and 1q12 [525,526,527,528]. Jumping translocations involving 1q are hallmarks of some hematological and other malignancies and can involve interstitial telomeric sequences [524,529] (see Section 6). Segmental copy number variations are also a characteristic of 1q abnormalities in MM and are most often represented by tandem duplications of the 1q21 band. The proportion of these abnormalities and their copy number increases as the disease progresses toward its advanced stages associated with a poor prognosis [512,530,531]. It should be noted that it is the 1q21 amplification that has been classified as the most prognostically significant in the heterogeneous class of 1q abnormalities [530]. The 1q21 region contains several genes whose deregulation or amplification are thought to impact MM progression, including the top candidate, CKS1B, encoding the cyclin-dependent kinase regulatory subunit 1B [532] as well as several other genes: PSMD4 encoding the proteasome 26S subunit, IL6R encoding the interleukin 6 receptor, MCL1 encoding the BCL2 family anti-apoptotic protein, ANP32E encoding H2A.Z histone chaperone, ADAR1 encoding adenosine deaminase, and ILF2 encoding the interleukin enhancer-binding factor 2 [476,530,531,533]. The 1q21 gains are associated with resistance to bortezomib with no impact on the thalidomide-based regimen [534].
Detailed analysis of 13q deletions revealed several types of aberrations, including small interstitial deletions and gross chromosomal aberrations of the entire 13q. The 13q aberrations were detected in plasma cells of 21–45% of patients with MGUS, and their proportion increased up to 86% for MM and its advanced stages [509,535,536]. Deletion hotspots were at 13q14 where 13q14.11–q14.3 encompasses the minimal deleted region carrying dozens of genes, including the known tumor suppressor RB1 gene [476,536]. Similar to several other genes discussed above, biallelic inactivation of RB1 is frequently observed in MM. A complete defect of RB1 is associated with relapse of the disease and portends poor prognosis [536,537]. Another important region on the 13q is the 13q21.33 band, where DIS3 is localized. DIS3 inactivation is frequently associated with 13q deletion. Interestingly, DIS3 shows the highest frequency of biallelic inactivation among all significantly mutated genes in MM [86,476]. It should be mentioned that 13q abnormalities often coincide with other genomic rearrangements such as t(4;14); thus, their prognostic value is challenging [78]. For instance, DIS3 mutations are associated with IGH translocations t(4;14) and t(14;16) and are less frequently observed in hyperdiploid tumors [28,86,328].
Deletion of 17p results in a loss of the tumor suppressor gene TP53 located in 17p13, leading to its hemizygotization [476,538]. The occurrence of del(17p) is about 9.5–11% in newly diagnosed MM and increases up to 75% as the disease progresses to its later stages [402,456,513]. Mutations frequently inactivate the remaining copy of the TP53 in del(17p) MM tumors as the disease progresses, conferring a total defect [29,86,537,538,539]. The TP53/17p13 abnormalities resulting in biallelic inactivation of the TP53 have a clear negative impact on progression-free survival and overall survival, and, thus, are regarded as driver events for MM relapse [61,82,86,328,530,538,539]. The hemizygous 17p deletion, if not accompanied by the inactivation of the remaining copy of TP53, is less harmful and has been changed to non-prognostic, although a recent study still suggests an association with poor outcomes [515,530].
MYC translocations are a distinct group common in MM cases (up to 50%) and involve the 8q24 chromosome region. In most cases, these translocations result in MYC upregulation. MYC is a transcriptional regulator and a well-established oncogene whose deregulation is associated with many different malignancies, including B cell lymphomas [540] (see Section 1.2). MYC is one of the most frequently amplified protein-coding genes in cancer. MYC is located in the 8q24 band, a hotspot region for chromosomal rearrangements and SNPs associated with different types of cancer and a frequent site of viral integration [541,542,543]. MYC translocations are rarely found in MGUS or stable SMM and are mostly considered to be driver events in MGUS/SMM to MM progression [48,51,62,83,544,545,546]. Because of participation in late-progression events, they often associate with adverse prognosis [456,547,548]. Approximately one-third of 8q24 translocations in MM events involve immunoglobulin (IG) loci (IGH, IGK, IGL-immunoglobulin heavy chain, kappa, lambda genes) where MYC juxtaposes to powerful IG enhancers (Figure 4). [545,549,550]. In the rest of the cases, MYC is juxtaposed to other super-enhancers associated with genes NSMCE2, TXNDC5, FAM46C, FOXO3, IGJ, and PRDM1 [547,549,550,551]. In addition, MYC is a direct target of IRF4, which is linked to myeloma predisposition [346,552]. MYC rearrangements often coincide with other karyotypic abnormalities: t(14;16) or chromoplexy [550,551]. In addition, IGL-MYC translocations constituting approximately 10% of all genomic abnormalities found in MM often coincide with the hyperdiploid state [548].

5.2. Mechanisms of Genomic Instability at 14q32 and 8q24 Regions

The IGH locus (14q32.33) is localized at the very extreme terminus of chromosome 14 in the telomeric band. Breakpoints at 14q32.33 initiate a variety of different translocations associated with B cell malignancies and other cancers. Remarkably, some of the partner loci have telomeric (4p16; 6p25) or subtelomeric (16q23) locations [484]. This fact implicates that telomere integrity and interaction between telomeric sequences on different chromosomes might play an essential role in generating these rearrangements. Lymphoid-specific genome rearrangement processes V(D)J recombination and CSR are thought to be the main sources of translocations observed in MM (Figure 5). A proportion of translocation events in MM occurs early in B-cell development at the pro-B cell stage in bone marrow due to aberrant V(D)J recombination [553]. V(D)J recombination is a process leading to diversification of the antigen-binding regions of immunoglobulins and occurs via random recombination between variable (V), diversity (D), and joining (J) genes segments (Figure 5). The other (major) proportion of chromosomal aberrations involving the IGH locus in MM occurs in mature B cells during CSR, a process that is responsible for the isotype switching of the heavy immunoglobulin chains. CSR occurs in antigen-activated naïve B-cells soon after infection or immunization (Figure 3) and results in substitution of the constant regions encoding IgM and IgD for the IgG, IgE, or IgA (Figure 5). Both V(D)J recombination and CSR result in intrachromosomal rearrangements, leading to deletion of the genetic information at the IGH locus, and are irreversible. In the case of CSR, the sequence encoding the constant CH μ (for IgM) region is substituted for another one encoding for a constant downstream region. The recombination targets during this process are switch (S) regions preceding every constant region except IGHD encoding the CH δ (Figure 5). S regions are essential elements providing CSR; they span 2–10 kb and are highly enriched in repeats and G-quadruplex-forming elements necessary for their function. The other elements at the human IGH locus are two super-enhancers, the 3′ regulatory regions (3′RRs) and the Eμ intronic enhancer, which are the master regulators of CSR, V(D)J recombination, and SHM and play an essential role in B-cell development.
CH genes in the human genome are arranged into two tandem clusters of genes encoding γ, ε and α isotypes, where Eμ intronic enhancer precedes the IGHM gene and 3′RRs are positioned downstream of each cluster (Figure 5) [578,579,580]. Specifically, 3′RRs are located ∼160 kb and ∼275 kb downstream of the Eμ intronic enhancer. Each 3′RRs in humans contains three enhancer elements, hs3, hs1-2, and hs4, which carry binding sites for regulatory and transcription factors [581,582,583]. The 3′RRs are enriched in repetitive sequences, G-quadruplex-forming elements, and carry inverted repeats flanking the hs1-2 regulatory element that makes the structure inside the 3′RR a quasi-palindrome [80,583,584]. The 3′RR and functional S-regions are sufficient to recruit AID and provide CSR when placed at an artificial location in the genome at the IGK locus [585]. Defects in the IGH regulatory enhancers in mice affect recombination at the IGH locus to varying degrees and may lead to a deficiency in antibody production [579,586,587,588].
CSR requires an action of AID and depends on the synthesis of noncoding RNA across the S regions. It was proposed that such transcription would result in R-loop formation and consequent recruitment of AID to the single-stranded regions [564,589,590]. Deamination of cytosine residues generates uracil in the targeted S-region. Subsequent uracil excision from the DNA triggers formation of DNA breaks that are thought to promote recombination. Partner S regions may be located as far as ~270 kb away from each other, so the efficient recombination requires their juxtaposition through long-range chromatin interaction. S/S synaptosome formation is a prerequisite for CSR: Two distant S regions are brought into proximity to facilitate rearrangement. Studies in mice suggest that the IGH locus undergoes a remarkable conformational change during the B cell differentiation, including the formation of chromatin loops that provide interaction between distant sequences and play an essential role in the IGH locus rearrangements. For instance, chromosome looping allows Vh segments to contact Dh and Jh segments during V(D)J recombination [591]. The contraction of the 2.8 Mb murine IGH region depends on WAPL cohesion-release factor, which is controlled by PAX5. Reduced Wapl expression causes global alterations in the chromosome architecture [592,593]. Another round of conformational changes occurs before CSR in mature B-cells. The Eμ intron enhancer and the 3′RR get positioned near each other with the formation of the loop [594]. This event is accompanied by a remarkable change in DNA methylation and histone acetylation in the 3′RR [580,595]. Upstream of each acceptor S region is GLT promoter, which, upon B cell activation with cytokines, can be recruited to the Eμ/3′RR. This event enables the effective transcription of the S regions targeted for CSR and their juxtaposition [590,594,596,597]. Cohesin complex and CTCF are implicated in mediating long-range interactions at the IGH locus: They tether bases of loops, regulate loop extrusion, and demarcate topologically associating domains [595,598]. The IGH locus in both human and murine genomes contain a remarkable number of CTCF sites; some of them are proposed to regulate long-range chromatin interaction, germline transcription, and CSR [598,599,600,601].
Intriguingly, functions of many genes recurrently mutated or deregulated in MM or associated with MM development can be linked to the regulation of the IGH locus transcription, V(D)J recombination, and CSR. Therefore, the processes at the IGH locus during B cell development can be the key to unraveling the origins of MM (Figure 6).
PAX5 is a major transcription regulator involved in B-cell development, and its mutational inactivation or deregulation is associated with B-cell malignancies [615,616,617]. Mutations in the cis-regulatory elements affecting PAX5 expression and structural rearrangements involving PAX5 were identified in MM [60,398,618,619]. Interestingly, PAX5 regulatory regions, as well regulatory regions of some other B-cell specific oncogenes, get in close proximity and undergo antisense transcription and are targeted by AID, which may explain translocations between these loci and the IGH locus and mutagenesis in these regions [614,620]. Differences in PAX5 expression profiles in B-cells between healthy people and MM patients and unique PAX5 isoforms have been noted in MM [621]. PAX5 has multiple sites of binding at the IGH locus and participates in spatial locus contraction necessary for its rearrangement and V-DJ recombination [622,623,624,625,626]. The 3′RRs in mice are interspersed with multiple PAX5 binding sites and undergo dynamic changes in PAX5-binding during CSR [596]. PAX5, together with Linker histone H1 is involved in B-cell specific DNA methylation and histone modifications at the 3’RR in mice [604]. DNA is hypermethylated at B-cell specific intronic enhancer regions in MM but not in normal plasma cells, and these sites overlap with binding sites of some transcription factors, including PAX5 [428]. PAX5 binding to the DNA recruits PTIP and promotes binding of MLL3/MLL4 histone lysine methyltransferase complex to specific regions at the IGH locus in mice, which promotes chromatin changes and transcription initiation at switch-regions critical for CSR [602,603]. MLL3/4 containing COMPASS complex includes PTIP and KDM6A and functions with p300/CBP in the de novo “commissioning” of enhancers [611]. As was discussed in Section 3, mutations in MLL3, MLL4, KDM6A, EP300, CBP, and, to a lesser extent, PTIP were all found in MM.
Interestingly, the IGH locus is one of the most heavily transcribed in plasma cells. Antisense transcription has been observed from variable region exons and from S regions and is thought to be implicated in AID-targeting [608,614]. SWI/SNF chromatin remodeling complexes are recruited to IGH locus and facilitate antisense IGH transcription and accessibility of IGH for recombination in B cell precursors [609]. As was discussed earlier (Section 2 and Section 3), genes encoding SWI/SNF complexes are altered in many cancers, including MM, and mutations in the genes encoding components of SWI/SNF complex are linked to MM predisposition. In hematopoietic progenitors and non-B lineage cells, IGH is localized to the nuclear periphery [627]. More specifically, in fibroblasts, the IGH locus was shown to be tethered to the nuclear lamina, where its transcription is repressed [628]. Relocalization of the IGH locus from the periphery to the nuclear interior occurs in pro-B cells that, along with intergenic and antisense transcription and action of chromatin remodelers, facilitates genes shuffling during V(D)J recombination and further maturation of the antigenic repertoire [627,629,630]. It has been proposed that in cells, there is a special perinucleolar recombination compartment where the IGH locus undergoes rearrangements during B-cell maturation and where the activity of AID and RAG1/2 can be sequestered [631].
AID recruitment to the transcribed S-regions at the IGH locus is aided by the exosome. Exosome interacts with the RNA Polymerase II and may facilitate exonucleolytic degradation of a 3′ free RNA end on backtracked complexes, thus helping to generate ssDNA substrate for AID. The DIS3 gene that we discussed in Section 2 and Section 3 is a catalytic component of the exosome and is implicated in regulating of AID-dependent mutagenesis and recombination at the IGH locus [331,332]. Other genes involved in transcription regulation, RNA processing, and maturation might also impact the complex process of the IGH locus reorganization. More research is required to understand the interrelation between various processes at the IGH locus and MM development.
Similar to IGH translocations, some types of MYC translocations are thought to originate from the aberrant, B-cell specific processes: V(D)J recombination, IGH switch recombination, and SHM [632]. Studies in a murine model with AID-deficient mice have suggested that AID plays a crucial role in MYC/IGH translocations [633]. However, the localization of MYC translocations breakpoints in the IGH locus in newly diagnosed MM patients showed that they are spread out over the constant regions with no apparent association with AID motif clusters, pointing to a different mechanism of their generation [551]. 8q24 is a very peculiar region – it is a so-called “gene desert” apparently representing an important source of genomic instability and translocation events. This region is almost devoid of protein-coding genes but contains multiple regulatory elements, and it is a source of noncoding DNA transcription [542]. Remarkably, the breakpoints at 8q24 found in MM surround a region of active chromatin characterized by H3K27Ac, H3K36me3, and H3K4me1 marks, as well as DNaseI hypersensitivity [551]. Human Plasmacytoma Variant Translocation 1 (PVT1) gene is located in the same chromosome band 8q24 just 54Kb away from MYC, and is frequently involved in structural rearrangements observed in MM, primarily in MYC rearrangements [634]. PVT1 encodes 52 noncoding RNAs, including circular ncRNAs, linear ncRNAs, and microRNAs functioning as negative regulators of immune response and associating with hematological malignancies and other types of cancer [635]. 8q24 translocations result in overexpression of MYC as well as PVT1 [551]. Besides that, the PVT1 promoter can inhibit MYC expression by outcompeting intragenic enhancers located in the PVT1 locus [636]. Interestingly, in mice, MYC is recruited to the same transcription factory as IGH, which might facilitate spatial closure of these regions and consequently specific chromosomal translocations involving IGH and MYC [637]. MYC and IGH loci in mice have been shown to interact physically with each other and are tethered to the nucleolus via the nucleolar organizing regions at their chromosomes [638,639]. Colocalization of IGH, MYC, and CCND1 loci at a time when B cells undergo V(D)J recombination or SHM/CSR might favor certain types of translocations [640,641].

5.3. Complex Chromosomal Rearrangements in MM and Their Mechanisms

Several types of complex chromosomal rearrangements have been observed in MM. They include chromoplexy, templated insertions, and chromothripsis [61,83,619,642,643,644]. Chromoplexy represents a complex genomic alteration leading to balanced rearrangements (i.e., without apparent cytogenetically detectable loss of genetic material) at multiple breakpoints of several chromosomes that are believed to occur in a coordinated fashion [645,646]. In chromoplexy, the rearrangements often occur as nearly precise junctions without large deletions; however, small deletions may occur involving those affecting function of oncosuppressors [647]. NHEJ is thought to be a primary mechanism of the generation of such rearrangements. Chromoplexy is considered a late-progression event in MM, possibly linked to positive selection and drug resistance development [61].
Templated insertions are a distinct class of structural rearrangements characterized by DNA segments copied from different areas of the genome, joined and inserted as one contiguous sequence into a single derivative chromosome [645,648]. These insertions are found in about 20% of MM cases and are clonal events, suggesting their early origin in MM development [61,83,619]. Templated insertions in MM can occur in one or several steps and lead to amplification of oncogenes (e.g., MYC) or their juxtaposition to powerful enhancers [61,547,619,645]. The term “chromoanasynthesis” is often used in literature to describe a replication-based process leading to complex rearrangements that are templated and characterized by gain or amplification of chromosomal segments [647]. Break-induced replication (BIR) is a specific HR mechanism responsible for the repair of DNA strand breaks independently of each other that involves template switching and replication, and that can account for complex rearrangements seen in cancers [649,650]. BIR can involve multiple template switching events between dispersed repeated sequences [651]. The two related processes termed microhomology-mediated, break-induced replication (MM-BIR) and fork-stalling and template switching (FoSTeS) are discussed to underlie complex genomic rearrangements that resemble the pattern of templated insertions observed in cancer cells [652,653,654,655].
Chromothripsis is a massive ‘‘catastrophic’’ structural rearrangement of the genome that involves the shattering of one or several chromosomes to pieces followed by their reassembly in a new order with the formation of new chromosomes. Chromothripsis is believed to occur in one step and, besides gene order changes, can result in deletions or amplification of certain regions. Chromothripsis also causes extrachromosomal structures, termed double minutes, containing pieces of shattered chromosomes with altered chromatin structure. Initially, chromothripsis was thought to be a rare event in cancer progression, but recent studies reveal that it is not. Chromothripsis occurs with a frequency of up to 30% in osteosarcoma and glioblastoma, about 60% in melanoma, and up to 36% in MM [61,619,647,648,656]. Chromothripsis, as well as templated insertions, are currently viewed as relatively early events in MM development that can drive tumor development [61,62]. Chromothripsis is associated with known high-risk genetic alterations in MM, including IGH translocations involving MMSET, MAF, or MAFB; biallelic inactivation of TP53, deletion of 1p12, and high APOBEC mutational burden [619]. In addition to that, chromothripsis was associated with poor outcomes and was linked to treatment in MM [619,642,644,657]. For instance, chromothripsis occurs more often in patients with bortezomib-resistant MM that can be interpreted as a drug-induced response [657].
Several mechanisms have been proposed to explain the nature of chromothripsis. The initiating events are thought to be telomeres’ attrition and dysfunction leading to telomere-telomere fusions and consequent breakage-fusion-bridge (BFB) cycles [656,658,659,660,661]. The dicentric chromosomes formed during this process are subject to processing aimed at their isolation or resolution during cell division. In one scenario, the dicentric chromosomes or chromosomes’ fragments can be compartmentalized into micronuclei. Such chromosomes become physically isolated from the rest of the genome in a microenvironment where they can undergo abnormal repair processes resulting in chromosomes’ shattering [662,663,664,665]. Micronuclei are often found in cancer cells and are characterized by abnormalities in DNA replication, transcription, and nuclear envelope structure [666]. Formation of micronuclei and chromosome pulverization distal to 1q12 have also been described in MM patients with aberrations involving 1q [526,644]. Another model proposes that chromatin bridges formed during cytokinesis are pulverized via concerted action of TREX1 exonuclease and APOBEC3A/B cytidine deaminases [658,667,668]. TREX1 generates ssDNA that is deaminated by APOBEC3A/B enzymes with the consequent repair of abasic sites leading to DNA breaks. This mechanism implies that chromothripsis and kataegis events can be interrelated. In some cancers, chromothripsis and other structural variations indeed associate with kataegis [452,648,669,670]. Remarkably, TREX1 is a cytoplasmic nuclease requiring nuclear envelope rupture to access the chromosomal DNA [658]. Other models suggest that multi-invasion-induced rearrangements (MIR) and MM-BIR can contribute to complex structural rearrangements observed in chromothripsis [668,671].

6. Telomere Maintenance Pathways and MM Risk

Multiple lines of evidence suggest that a major risk of developing MM can be linked to defects in telomere maintenance mechanism leading to changes in telomere length and structure. Telomere shortening has been observed in patients with plasma cell disorders [672]. MM tumor cells can maintain stable short telomeres and telomere length is a prognostic marker in MM [673,674]. Telomerase activity negatively correlates with MM survival rate and telomere length in MM, according to telomerase maintaining of short and critically short telomeres [675,676,677]. At the same time, a small proportion of MM patients have very long telomeres alluding to Alternative Lengthening of Telomeres (ALT) in the progress of the disease.
A strong correlation was found between telomerase activity and 1q jumping translocations [675], implicating telomerase-dependent DSB repair or targeted telomere insertion in the process of 1q rearrangements [529,678,679,680]. It is noteworthy that a block of Interstitial Telomeric Sequences (ITSs) was reported at 1q12 [681]. ITSs are important hotspots of DSBs and recombination and have been implicated in the formation of jumping translocations [529,682,683,684]. ITSs can also impact 3D nuclear architecture through interaction with telomeres and ITL loops mediated by TRF2 and Lamin A/C [685,686]. Changes in the 3D nuclear architecture accompanied by telomere attrition and formation of telomere aggregates have been noted during disease progression from MGUS to MM [687]. A number of telomeric signals were reported to be significantly elevated in MM compared to MGUS, indicating either addition of new chromosomes or amplification of interstitial telomeric repeats [688]. Other than telomerase activation, a significant increase of TRF2 and TANK1 expression levels in patients with MGUS and MM and TRF1 in patients with MM was observed [676]. Telomere dysfunction and attrition can promote recombination between telomeres, followed by telomere-telomere fusions and BFB cycles and eventually chromothripsis, as was discussed in Section 5.3.
Several SNPs potentially affecting genes implicated in telomere length regulation have been associated with the increased susceptibility to MM (Table S3). Some of these polymorphisms lie within regions occupied by TERC and TERT genes encoding for telomerase RNA and telomerase catalytic subunit, correspondingly. These SNPs may affect telomerase activity [343,689]. The risk of B-cell lymphomas, including MM, is associated with longer telomeres [690,691]. For instance, a common [C] allele of rs10936599 is strongly associated with increased telomere length and susceptibility to MM and colorectal cancer [343,689,691,692,693,694]. Although the association of this SNP with telomerase activity remains controversial, the carrier status of [C] allele for the rs10936599 was associated with increased expression of TERC in another tumor of differentiated B-cells, Chronic Lymphocytic Leukemia (CLL), and susceptibility to CLL [695]. Carriers of another allele [T] at rs10936599 demonstrated shorter telomeres compared to homozygous [C] carriers [696]. rs10936599 has a genetic association with SNP rs2293607, which maps 63bps 5′ to TERC and impacts TERC mRNA expression [343,692]. The rs2293607 [A] allele results in longer telomeres in colorectal cancer cell lines [692]. Another SNP, rs2242652, which lies in intron four of the TERT gene, is associated with a decreased risk of MM [689]. The minor allele [A] of this SNP downregulates TERT promoter activity and has been associated with the increased breast and ovarian cancer but decreased risk of prostate cancer and MM [689,697]. Additionally, SNP rs10936600 is located in the same 3q26.2 locus and is associated with an increased risk of MM [354]. This SNP is located within the LRRC34 gene—a predicted ribonuclease inhibitor. Interestingly, the telomere-associated SNP rs10936599 affected LRRC34 expression, and the risk allele for shorter telomere length [T] was associated with reduced LRRC34 expression [698]. Another SNP, rs58618031, associated with increased risk of MM, is located in the band 7q31.33, which lies within the POT1-AS1 lncRNA gene near the POT1 gene and might affect POT1 expression [345]. POT1 is a component of the Shelterin complex protecting telomeres by interaction with the single-stranded G-rich tip of a telomeric DNA [699]. POT1 mutations are commonly observed in some B-cell malignancies such as CLL [700]. Increased POT1 expression was linked to the transformation from MGUS to MM [701]. In addition to that, differences in the expression of other telomere maintenance genes between MGUS and MM were reported [673,676]. RECQL, encoding RECQ1 helicase associating with telomeres in ALT cells [702] is significantly overexpressed in MM, which correlates with resistance to melphalan and bortezomib [703,704]. RECQL overexpression in MM is linked to aberrant methylation of miR-203 [703].
Genes LSD1/KDM1A, ARID1A, USP45, DIS3, and EP300, where family risk mutations for MM have been reported (see Section 2), are also linked to the telomere maintenance pathway [305,319,330]. LSD1 interacts with telomeric repeat containing RNA TERRA and regulates silencing at telomeres and telomere length [314,705,706]. ARID1A represses TERT transcription and promotes telomere cohesion and protection [489,707,708]. USP45 controls the activity of the ERCC1-XPF complex that interacts with Shelterin component TRF2 and facilitates nucleolytic processing of the 3′-end at uncapped telomeres, promotes t-loop formation, and plays a role in telomere integrity [325,709]. DIS3 is implicated in the processing of a variety of RNA transcripts, including telomerase RNA (hTR) precursors, which can affect telomerase activity [710]. As was discussed above, RNA exosome and DIS3 have been implicated in mutagenesis and recombination in the IGH locus as well, specifically in AID targeting during CSR [331,332,711]. DIS3 can potentially play a dual role in MM, affecting both telomerase activity and recombination at the IGH locus, where primary structural aberrations during the development of MM occur. p300 acetylates TRF2, and this modification stabilizes the TRF2 protein by inhibiting its ubiquitin-dependent proteolysis and promotes efficient binding of TRF2 to telomeres [712]. p300 also associates with telomeres and catalyzes acetylation of H3K27 [713]. It is necessary to re-emphasize, that the position of the IGH locus itself in the telomeric region could be an important factor influencing its chromatin structure, 3D position in the nucleus, and the DNA repair that can ultimately forge structural aberrations at the IGH locus. In conclusion, telomere maintenance pathways play a significant and multifaceted role in the development of MM. Telomerase might be the general target of the PIs, such as bortezomib, carfilzomib, epoxomicin, and others, affecting its activity both transcriptionally and post-translationally [714,715].

7. Conclusions and Further Perspectives

We have reviewed the current understanding of the genetic heterogeneity and variability of MM and discussed different aspects of this phenomenon: types of genetic alterations found in MM, the molecular mechanisms of MM genome plasticity, as well as the dynamics of genetic changes during the development of the disease and its treatment. Owing to extensive research of MM and the wide use of NGS, the molecular mechanisms of MM development are much better understood now. Dozens of recurrently mutated genes were found in MM genomes. NGS also contributed to a better understanding of predisposition to MM and helped in a precise analysis of structural genome variations seen in MM. Recent WES studies revealed genomic signatures that may be useful in molecular individualizing of B-cell neoplasms and a better understanding their origin. For instance, the mutational landscapes of MM, plasmablastic lymphoma (PBL), and diffuse large B-cell lymphoma (DLBCL) differ in the proportion of mutations affecting the signaling pathways RAS/MAPK, JAK/STAT, and NOTCH. Another remarkable feature is different frequencies of mutations affecting DIS3 and TET2 genes [716,717,718].
Over the past few years, the overall survival of patients with MM has been significantly increased. This improvement is made possible by the development of new drugs and patients’ stratification based on genetic factors. Despite this undoubted success, the disease remains incurable and relapses in most patients after a certain period. Moreover, some patients do not ever respond to any currently available therapy. Several questions have to be answered. It is vital to find why some patients do not respond to therapies or develop drug resistance. What is the effect of the treatment on tumor evolution? This problem becomes even more important nowadays because new drugs for MM treatment appear every year. Some of these drugs are based on new principles of action (e.g., monoclonal antibodies, CAR T-cell therapy, targeted delivery of chemotherapeutic drugs, such as melphalan flufenamide or antibody–drug conjugates), while others are developed on a platform of new knowledge about the genetic plasticity of the MM genome (e.g., histone deacetylase inhibitors, drugs targeting DNA or histone’s methylation pathways, pyrimidine nucleoside analogs) [103,130,132,719,720,721]. Next-generation IMiDs, PIs, agents that target the ubiquitin proteasomal cascade, and signaling pathways, are being developed to overcome the limitations of existing therapies [719,722,723].
While genome instability has been acknowledged as the driving force in MM development, we still do not know much about the combined contribution of the intrinsic genetic plasticity of MM and therapy to the progression of the disease [724]. Epigenetic factors and non-coding RNome add another level of complexity [725,726].
An urgent problem is a search for new risk factors and more fractional patient stratification based on certain risk factors that will allow using highly personalized regiments. Though the NGS techniques led to progress in understanding pathogenesis and progression of MM, WES and WGS are applied in most cases for basic research but not as a diagnostic technique that would help assess the severity of the condition and response to treatment. The WGS approach has cogent value, but bioinformatics analysis of NGS data for many samples is time-consuming and is not widely used in routine clinical care. The cytogenetic analysis by fluorescence in situ hybridization (FISH), karyotyping, or single nucleotide polymorphism arrays are prevalent approaches in current clinical practice for MM analysis. Custom target NGS has significant advantages to identify mutations, copy number alterations, and translocations in clinical practice [328,554,727,728]. The ultra-low pass whole-genome sequencing (ULP-WGS) approach can be an alternative method to WGS for the detection of copy number aberrations [48,437,439,729]. The DNA sequence data in combination with the whole transcriptome approach (RNA-Seq) helps to find driver mutations, molecular heterogeneity in different stages of MM, assess drugs response, and guide therapeutic decisions [730,731,732]. Newer studies attempted to create a platform for precision treatment and extend the application of WGS and RNA-Seq approach for MM in clinics [80,730,732], but the low level of standardization of this technique so far precludes its routine use in clinical practice. In the future, MM research utilizing comprehensive genomic data in combination with clinical evidence will robustly identify genetic markers associated with MM and eventually lead to improved diagnostics and treatment of MM.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/cancers13235949/s1. Table S1: NGS-based studies of MM genomes, Table S2: Recurrently mutated genes in MM, Table S3: SNPs associated with MM/MGUS according to GWAS and other studies.

Author Contributions

Conceptualization, first draft, and final revision, A.Y.A. Analysis of published whole genome sequencing and whole exome sequencing data, A.S.Z., A.G.L. and A.Y.A. Figures and tables preparation, A.Y.A., A.G.L., A.S.Z., I.V.Z., E.I.S., I.I.K. and S.V.G. Supervision, A.Y.A. and Y.I.P. All authors participated in writing of certain sections and reviewed the whole manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Russian Science Foundation (RSF), grant # 20-15-00081.

Acknowledgments

Computational analysis of breakpoints’ location with genomic elements on the human IGH locus was supported by the Government of the Russian Federation through the ITMO Fellowship and Professorship Program to A.S.Z. The authors also thank Research Park of the St. Petersburg State University for data analysis support. The authors acknowledge the support from the St. Petersburg State University (project 73024371) and the Center for Molecular and Cell Technologies (Research Park, St. Petersburg State University) to A.Y.A. The authors thank Sarah Holstein for critical reading of the manuscript. The authors also thank Jeffrey Patterson for expert English editing. Icons in preparation of some figures were adapted from BioRender.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rajkumar, S.V. Multiple myeloma: 2020 update on diagnosis, risk-stratification and management. Am. J. Hematol. 2020, 95, 548–567. [Google Scholar] [CrossRef] [PubMed]
  2. Rajkumar, S.V. Updated Diagnostic Criteria and Staging System for Multiple Myeloma. Am. Soc. Clin. Oncol. Educ. Book 2016, 35, e418–e423. [Google Scholar] [CrossRef] [PubMed]
  3. Rajkumar, S.V.; Dimopoulos, M.A.; Palumbo, A.; Blade, J.; Merlini, G.; Mateos, M.V.; Kumar, S.; Hillengass, J.; Kastritis, E.; Richardson, P.; et al. International Myeloma Working Group updated criteria for the diagnosis of multiple myeloma. Lancet Oncol. 2014, 15, e538–e548. [Google Scholar] [CrossRef]
  4. Kyle, R. Monoclonal gammopathy of undetermined significance. Natural history in 241 cases. Am. J. Med. 1978, 64, 814–826. [Google Scholar] [CrossRef]
  5. Dhodapkar, M.V. MGUS to myeloma: A mysterious gammopathy of underexplored significance. Blood 2016, 128, 2599–2606. [Google Scholar] [CrossRef]
  6. Dispenzieri, A.; Katzmann, J.A.; Kyle, R.A.; Larson, D.R.; Melton, L.J.; Colby, C.L.; Therneau, T.M.; Clark, R.; Kumar, S.K.; Bradwell, A.; et al. Prevalence and risk of progression of light-chain monoclonal gammopathy of undetermined significance: A retrospective population-based cohort study. Lancet 2010, 375, 1721–1728. [Google Scholar] [CrossRef]
  7. Kyle, R.A.; Therneau, T.M.; Rajkumar, S.V.; Larson, D.R.; Plevak, M.F.; Offord, J.R.; Dispenzieri, A.; Katzmann, J.A.; Melton, L.J. Prevalence of Monoclonal Gammopathy of Undetermined Significance. N. Engl. J. Med. 2006, 354, 1362–1369. [Google Scholar] [CrossRef] [PubMed]
  8. Chudasama, R.; Barth, P. Risk Stratification of Precursors to Multiple Myeloma in 2020. Rhode Isl. Med. J. 2020, 103, 46–47. [Google Scholar]
  9. Kyle, R.A.; Therneau, T.M.; Rajkumar, S.V.; Offord, J.R.; Larson, D.R.; Plevak, M.F.; Melton, L.J. A Long-Term Study of Prognosis in Monoclonal Gammopathy of Undetermined Significance. N. Engl. J. Med. 2002, 346, 564–569. [Google Scholar] [CrossRef] [PubMed]
  10. Kyle, R.A.; Larson, D.R.; Therneau, T.M.; Dispenzieri, A.; Kumar, S.; Cerhan, J.R.; Rajkumar, S.V. Long-Term Follow-up of Monoclonal Gammopathy of Undetermined Significance. N. Engl. J. Med. 2018, 378, 241–249. [Google Scholar] [CrossRef]
  11. Landgren, O.; Kyle, R.A.; Pfeiffer, R.M.; Katzmann, J.A.; Caporaso, N.E.; Hayes, R.B.; Dispenzieri, A.; Kumar, S.; Clark, R.J.; Baris, D.; et al. Monoclonal gammopathy of undetermined significance (MGUS) consistently precedes multiple myeloma: A prospective study. Blood 2009, 113, 5412–5417. [Google Scholar] [CrossRef] [PubMed]
  12. Kyle, R.; Greipp, P. Smoldering multiple myeloma. N. Engl. J. Med. 1980, 302, 1347–1349. [Google Scholar] [CrossRef]
  13. Kyle, R.A.; Remstein, E.D.; Therneau, T.M.; Dispenzieri, A.; Kurtin, P.J.; Hodnefield, J.M.; Larson, D.R.; Plevak, M.F.; Jelinek, D.F.; Fonseca, R.; et al. Clinical Course and Prognosis of Smoldering (Asymptomatic) Multiple Myeloma. N. Engl. J. Med. 2007, 356, 2582–2590. [Google Scholar] [CrossRef]
  14. Weiss, B.M.; Abadie, J.; Verma, P.; Howard, R.S.; Kuehl, W.M. A monoclonal gammopathy precedes multiple myeloma in most patients. Blood 2009, 113, 5418–5422. [Google Scholar] [CrossRef]
  15. Lonial, S.; Dhodapkar, M.V.; Rajkumar, S.V. Smoldering myeloma and the art of war. J. Clin. Oncol. 2020, 38, 2363–2365. [Google Scholar] [CrossRef] [PubMed]
  16. Kyle, R.A.; Dispenzieri, A.; Kumar, S.; Larson, D.; Therneau, T.; Vincent Rajkumar, S. IgM monoclonal gammopathy of undetermined significance (MGUS) and smoldering Waldenström’s macroglobulinemia (SWM). Clin. Lymphoma Myeloma Leuk. 2011, 11, 74–76. [Google Scholar] [CrossRef] [PubMed]
  17. Girard, L.P.; Soekojo, C.Y.; Ooi, M.; Poon, L.M.; Chng, W.J.; de Mel, S. Immunoglobulin M paraproteinaemias. Cancers 2020, 12, 1688. [Google Scholar] [CrossRef]
  18. Kyle, R.A.; Gertz, M.A.; Witzig, T.E.; Lust, J.A.; Lacy, M.Q.; Dispenzieri, A.; Fonseca, R.; Rajkumar, S.V.; Offord, J.R.; Larson, D.R.; et al. Review of 1027 patients with newly diagnosed multiple myeloma. Mayo Clin. Proc. 2003, 78, 21–33. [Google Scholar] [CrossRef]
  19. Chawla, S.S.; Kumar, S.K.; Dispenzieri, A.; Greenberg, A.J.; Larson, D.R.; Kyle, R.A.; Lacy, M.Q.; Gertz, M.A.; Rajkumar, S.V. Clinical course and prognosis of non-secretory multiple myeloma. Eur. J. Haematol. 2015, 95, 57–64. [Google Scholar] [CrossRef] [PubMed]
  20. Dutta, A.K.; Fink, J.L.; Grady, J.P.; Morgan, G.J.; Mullighan, C.G.; To, L.B.; Hewett, D.R.; Zannettino, A.C.W. Subclonal evolution in disease progression from MGUS/SMM to multiple myeloma is characterised by clonal stability. Leukemia 2018, 33, 457–468. [Google Scholar] [CrossRef] [PubMed]
  21. Rasche, L.; Chavan, S.S.; Stephens, O.W.; Patel, P.; Tytarenko, R.; Ashby, C.; Bauer, M.; Stein, C.; Deshpande, S.; Wardell, C.; et al. Spatial genomic heterogeneity in multiple myeloma revealed by multi-region sequencing. Nat. Commun. 2017, 8, 268. [Google Scholar] [CrossRef] [PubMed]
  22. Keats, J.J.; Chesi, M.; Egan, J.B.; Garbitt, V.M.; Palmer, S.E.; Braggio, E.; Van Wier, S.; Blackburn, P.R.; Baker, A.S.; Dispenzieri, A.; et al. Clonal competition with alternating dominance in multiple myeloma. Blood 2012, 120, 1067–1076. [Google Scholar] [CrossRef]
  23. Walker, B.A.; Wardell, C.P.; Melchor, L.; Brioli, A.; Johnson, D.C.; Kaiser, M.F.; Mirabella, F.; Lopez-Corral, L.; Humphray, S.; Murray, L.; et al. Intraclonal heterogeneity is a critical early event in the development of myeloma and precedes the development of clinical symptoms. Leukemia 2014, 28, 384–390. [Google Scholar] [CrossRef]
  24. Kortüm, K.M.; Langer, C.; Monge, J.; Bruins, L.; Zhu, Y.X.; Shi, C.X.; Jedlowski, P.; Egan, J.B.; Ojha, J.; Bullinger, L.; et al. Longitudinal analysis of 25 sequential sample-pairs using a custom multiple myeloma mutation sequencing panel (M3P). Ann. Hematol. 2015, 94, 1205–1211. [Google Scholar] [CrossRef] [PubMed]
  25. Weinhold, N.; Ashby, C.; Rasche, L.; Chavan, S.S.; Stein, C.; Stephens, O.W.; Tytarenko, R.; Bauer, M.A.; Meissner, T.; Deshpande, S.; et al. Clonal selection and double-hit events involving tumor suppressor genes underlie relapse in myeloma. Blood 2016, 128, 1735–1744. [Google Scholar] [CrossRef]
  26. Egan, J.B.; Shi, C.X.; Tembe, W.; Christoforides, A.; Kurdoglu, A.; Sinari, S.; Middha, S.; Asmann, Y.; Schmidt, J.; Braggio, E.; et al. Whole-genome sequencing of multiple myeloma from diagnosis to plasma cell leukemia reveals genomic initiating events, evolution, and clonal tides. Blood 2012, 120, 1060–1066. [Google Scholar] [CrossRef]
  27. Melchor, L.; Brioli, A.; Wardell, C.P.; Murison, A.; Potter, N.E.; Kaiser, M.F.; Fryer, R.A.; Johnson, D.C.; Begum, D.B.; Wilson, S.H.; et al. Single-cell genetic analysis reveals the composition of initiating clones and phylogenetic patterns of branching and parallel evolution in myeloma. Leukemia 2014, 28, 1705–1715. [Google Scholar] [CrossRef] [PubMed]
  28. Lohr, J.G.; Stojanov, P.; Carter, S.L.; Cruz-Gordillo, P.; Lawrence, M.S.; Auclair, D.; Sougnez, C.; Knoechel, B.; Gould, J.; Saksena, G.; et al. Widespread genetic heterogeneity in multiple myeloma: Implications for targeted therapy. Cancer Cell 2014, 25, 91–101. [Google Scholar] [CrossRef]
  29. Bolli, N.; Avet-Loiseau, H.; Wedge, D.C.; Van Loo, P.; Alexandrov, L.B.; Martincorena, I.; Dawson, K.J.; Iorio, F.; Nik-Zainal, S.; Bignell, G.R.; et al. Heterogeneity of genomic evolution and mutational profiles in multiple myeloma. Nat. Commun. 2014, 5, 1–13. [Google Scholar] [CrossRef]
  30. Ledergor, G.; Weiner, A.; Zada, M.; Wang, S.Y.; Cohen, Y.C.; Gatt, M.E.; Snir, N.; Magen, H.; Koren-Michowitz, M.; Herzog-Tzarfati, K.; et al. Single cell dissection of plasma cell heterogeneity in symptomatic and asymptomatic myeloma. Nat. Med. 2018, 24, 1867–1876. [Google Scholar] [CrossRef]
  31. Bochtler, T.; Merz, M.; Hielscher, T.; Granzow, M.; Hoffmann, K.; Krämer, A.; Raab, M.S.; Hillengass, J.; Seckinger, A.; Kimmich, C.; et al. Cytogenetic intraclonal heterogeneity of plasma cell dyscrasia in AL amyloidosis as compared with multiple myeloma. Blood Adv. 2018, 2, 2607–2618. [Google Scholar] [CrossRef] [PubMed]
  32. Paíno, T.; Paiva, B.; Sayagués, J.M.; Mota, I.; Carvalheiro, T.; Corchete, L.A.; Aires-Mejía, I.; Pérez, J.J.; Sanchez, M.L.; Barcena, P.; et al. Phenotypic identification of subclones in multiple myeloma with different chemoresistant, cytogenetic and clonogenic potential. Leukemia 2015, 29, 1186–1194. [Google Scholar] [CrossRef] [PubMed]
  33. Morgan, G.J.; Walker, B.A.; Davies, F.E. The genetic architecture of multiple myeloma. Nat. Rev. Cancer 2012, 12, 335–348. [Google Scholar] [CrossRef] [PubMed]
  34. Walker, B.A.; Wardell, C.P.; Melchor, L.; Hulkki, S.; Potter, N.E.; Johnson, D.; Fenwick, K.; Kozarewa, I.; Gonzalez, D.; Lord, C.; et al. Intraclonal heterogeneity and distinct molecular mechanisms characterize the development of t(4;14) and t(11;14) myeloma. Blood 2012, 120, 1077–1086. [Google Scholar] [CrossRef] [PubMed]
  35. Nowell, P.C. The clonal evolution of tumor cell populations. Science 1976, 194, 23–28. [Google Scholar] [CrossRef] [PubMed]
  36. Greaves, M.; Maley, C.C. Clonal evolution in cancer. Nature 2012, 481, 306–313. [Google Scholar] [CrossRef]
  37. Pon, J.R.; Marra, M.A. Driver and passenger mutations in cancer. Annu. Rev. Pathol. Mech. Dis. 2015, 10, 25–50. [Google Scholar] [CrossRef] [PubMed]
  38. Baylin, S.B.; Jones, P.A. Epigenetic determinants of cancer. Cold Spring Harb. Perspect. Biol. 2016, 8, a019505. [Google Scholar] [CrossRef] [PubMed]
  39. Venkatesan, S.; Swanton, C.; Taylor, B.S.; Costello, J.F. Treatment-Induced Mutagenesis and Selective Pressures Sculpt Cancer Evolution. Cold Spring Harb. Perspect. Med. 2017, 7, a026617. [Google Scholar] [CrossRef]
  40. Chatterjee, A.; Rodger, E.J.; Eccles, M.R. Epigenetic drivers of tumourigenesis and cancer metastasis. Semin. Cancer Biol. 2018, 51, 149–159. [Google Scholar] [CrossRef] [PubMed]
  41. Davis, A.; Gao, R.; Navin, N. Tumor evolution: Linear, branching, neutral or punctuated? Biochim. Biophys. Acta 2017, 1867, 151. [Google Scholar] [CrossRef]
  42. Furukawa, Y.; Kikuchi, J. Molecular basis of clonal evolution in multiple myeloma. Int. J. Hematol. 2020, 111, 496–511. [Google Scholar] [CrossRef] [PubMed]
  43. Bahlis, N.J. Darwinian evolution and tiding clones in multiple myeloma. Blood 2012, 120, 927–928. [Google Scholar] [CrossRef]
  44. Magrangeas, F.; Avet-Loiseau, H.; Gouraud, W.; Lodé, L.; Decaux, O.; Godmer, P.; Garderet, L.; Voillat, L.; Facon, T.; Stoppa, A.M.; et al. Minor clone provides a reservoir for relapse in multiple myeloma. Leukemia 2013, 27, 473–481. [Google Scholar] [CrossRef] [PubMed]
  45. Jones, J.R.; Weinhold, N.; Ashby, C.; Walker, B.A.; Wardell, C.; Pawlyn, C.; Rasche, L.; Melchor, L.; Cairns, D.A.; Gregory, W.M.; et al. Clonal evolution in myeloma: The impact of maintenance lenalidomide and depth of response on the genetics and sub-clonal structure of relapsed disease in uniformly treated newly diagnosed patients. Haematologica 2019, 104, 1440–1450. [Google Scholar] [CrossRef] [PubMed]
  46. Diamond, B.; Yellapantula, V.; Rustad, E.H.; Maclachlan, K.H.; Mayerhoefer, M.; Kaiser, M.; Morgan, G.; Landgren, O.; Maura, F. Positive selection as the unifying force for clonal evolution in multiple myeloma. Leukemia 2021, 35, 1511–1515. [Google Scholar] [CrossRef]
  47. Bolli, N.; Maura, F.; Minvielle, S.; Gloznik, D.; Szalat, R.; Fullam, A.; Martincorena, I.; Dawson, K.J.; Samur, M.K.; Zamora, J.; et al. Genomic patterns of progression in smoldering multiple myeloma. Nat. Commun. 2018, 9, 1–10. [Google Scholar] [CrossRef] [PubMed]
  48. Bustoros, M.; Sklavenitis-Pistofidis, R.; Park, J.; Redd, R.; Zhitomirsky, B.; Dunford, A.J.; Salem, K.; Tai, Y.T.; Anand, S.; Mouhieddine, T.H.; et al. Genomic profiling of smoldering multiple myeloma identifies patients at a high risk of disease progression. J. Clin. Oncol. 2020, 38, 2380–2389. [Google Scholar] [CrossRef]
  49. Johnson, D.C.; Lenive, O.; Mitchell, J.; Jackson, G.; Owen, R.; Drayson, M.; Cook, G.; Jones, J.R.; Pawlyn, C.; Davies, F.E.; et al. Neutral tumor evolution in myeloma is associated with poor prognosis. Blood 2017, 130, 1639–1643. [Google Scholar] [CrossRef]
  50. Merz, M.; Hielscher, T.; Schult, D.; Mai, E.K.; Raab, M.S.; Hillengass, J.; Seckinger, A.; Hose, D.; Granzow, M.; Jauch, A.; et al. Cytogenetic subclone formation and evolution in progressive smoldering multiple myeloma. Leukemia 2020, 34, 1192–1196. [Google Scholar] [CrossRef]
  51. Boyle, E.M.; Deshpande, S.; Tytarenko, R.; Ashby, C.; Wang, Y.; Bauer, M.A.; Johnson, S.K.; Wardell, C.P.; Thanendrarajan, S.; Zangari, M.; et al. The molecular make up of smoldering myeloma highlights the evolutionary pathways leading to multiple myeloma. Nat. Commun. 2021, 12, 293. [Google Scholar] [CrossRef]
  52. Mikulasova, A.; Wardell, C.P.; Murison, A.; Boyle, E.M.; Jackson, G.H.; Smetana, J.; Kufova, Z.; Pour, L.; Sandecka, V.; Almasi, M.; et al. The spectrum of somatic mutations in monoclonal gammopathy of undetermined significance indicates a less complex genomic landscape than that in multiple myeloma. Haematologica 2017, 102, 1617–1625. [Google Scholar] [CrossRef] [PubMed]
  53. Corre, J.; Cleynen, A.; Robiou du Pont, S.; Buisson, L.; Bolli, N.; Attal, M.; Munshi, N.; Avet-Loiseau, H. Multiple myeloma clonal evolution in homogeneously treated patients. Leukemia 2018, 32, 2636–2647. [Google Scholar] [CrossRef] [PubMed]
  54. Ziccheddu, B.; Biancon, G.; Bagnoli, F.; De Philippis, C.; Maura, F.; Rustad, E.H.; Dugo, M.; Devecchi, A.; De Cecco, L.; Sensi, M.; et al. Integrative analysis of the genomic and transcriptomic landscape of double-refractory multiple myeloma. Blood Adv. 2020, 4, 830–844. [Google Scholar] [CrossRef] [PubMed]
  55. López-Corral, L.; Gutiérrez, N.C.; Vidriales, M.B.; Mateos, M.V.; Rasillo, A.; García-Sanz, R.; Paiva, B.; San Miguel, J.F. The progression from MGUS to smoldering myeloma and eventually to multiple myeloma involves a clonal expansion of genetically abnormal plasma cells. Clin. Cancer Res. 2011, 17, 1692–1700. [Google Scholar] [CrossRef]
  56. Mailankody, S.; Kazandjian, D.; Korde, N.; Roschewski, M.; Manasanch, E.; Bhutani, M.; Tageja, N.; Kwok, M.; Zhang, Y.; Zingone, A.; et al. Baseline mutational patterns and sustained MRD negativity in patients with high-risk smoldering myeloma. Blood Adv. 2017, 1, 1911–1918. [Google Scholar] [CrossRef] [PubMed]
  57. Maura, F.; Bolli, N.; Rustad, E.H.; Hultcrantz, M.; Munshi, N.; Landgren, O. Moving from Cancer Burden to Cancer Genomics for Smoldering Myeloma: A Review. JAMA Oncol. 2020, 6, 425–432. [Google Scholar] [CrossRef] [PubMed]
  58. Samur, M.K.; Samur, A.A.; Fulciniti, M.; Szalat, R.; Han, T.; Shammas, M.; Richardson, P.; Magrangeas, F.; Minvielle, S.; Corre, J.; et al. Genome-wide somatic alterations in multiple myeloma reveal a superior outcome group. J. Clin. Oncol. 2020, 38, 3107–3118. [Google Scholar] [CrossRef]
  59. Samur, M.K.; Roncador, M.; Aktas-Samur, A.; Fulciniti, M.; Bazarbachi, A.H.; Szalat, R.; Shammas, M.A.; Sperling, A.S.; Richardson, P.G.; Magrangeas, F.; et al. High-Dose Melphalan Significantly Increases Mutational Burden in Multiple Myeloma Cells at Relapse: Results from a Randomized Study in Multiple Myeloma. Blood 2020, 136, 4–5. [Google Scholar] [CrossRef]
  60. Hoang, P.H.; Cornish, A.J.; Sherborne, A.L.; Chubb, D.; Kimber, S.; Jackson, G.; Morgan, G.J.; Cook, G.; Kinnersley, B.; Kaiser, M.; et al. An enhanced genetic model of relapsed IGH-translocated multiple myeloma evolutionary dynamics. Blood Cancer J. 2020, 10, 101. [Google Scholar] [CrossRef]
  61. Maura, F.; Bolli, N.; Angelopoulos, N.; Dawson, K.J.; Leongamornlert, D.; Martincorena, I.; Mitchell, T.J.; Fullam, A.; Gonzalez, S.; Szalat, R.; et al. Genomic landscape and chronological reconstruction of driver events in multiple myeloma. Nat. Commun. 2019, 10, 3835. [Google Scholar] [CrossRef]
  62. Oben, B.; Froyen, G.; Maclachlan, K.H.; Leongamornlert, D.; Abascal, F.; Zheng-Lin, B.; Yellapantula, V.; Derkach, A.; Geerdens, E.; Diamond, B.T.; et al. Whole-genome sequencing reveals progressive versus stable myeloma precursor conditions as two distinct entities. Nat. Commun. 2021, 12, 1861. [Google Scholar] [CrossRef] [PubMed]
  63. Rustad, E.H.; Yellapantula, V.; Leongamornlert, D.; Bolli, N.; Ledergor, G.; Nadeu, F.; Angelopoulos, N.; Dawson, K.J.; Mitchell, T.J.; Osborne, R.J.; et al. Timing the initiation of multiple myeloma. Nat. Commun. 2020, 11, 1917. [Google Scholar] [CrossRef] [PubMed]
  64. Poh, C.; Keegan, T.; Rosenberg, A.S. Second primary malignancies in multiple myeloma: A review. Blood Rev. 2021, 46, 100757. [Google Scholar] [CrossRef]
  65. Musto, P.; Anderson, K.C.; Attal, M.; Richardson, P.G.; Badros, A.; Hou, J.; Comenzo, R.; Du, J.; Durie, B.G.M.; San Miguel, J.; et al. Second primary malignancies in multiple myeloma: An overview and IMWG consensus. Ann. Oncol. 2017, 28, 228–245. [Google Scholar] [CrossRef]
  66. Maclachlan, K.; Diamond, B.; Maura, F.; Hillengass, J.; Turesson, I.; Landgren, C.O.; Kazandjian, D. Second malignancies in multiple myeloma; emerging patterns and future directions. Best Pract. Res. Clin. Haematol. 2020, 33, 101144. [Google Scholar] [CrossRef]
  67. Engelhardt, M.; Ihorst, G.; Landgren, O.; Pantic, M.; Reinhardt, H.; Waldschmidt, J.; May, A.M.; Schumacher, M.; Kleber, M.; Wäsch, R. Large registry analysis to accurately define second malignancy rates and risks in a well-characterized cohort of 744 consecutive multiple myeloma patients followed-up for 25 years. Haematologica 2015, 100, 1340–1349. [Google Scholar] [CrossRef] [PubMed]
  68. Govindarajan, R.; Jagannath, S.; Flick, J.T.; Vesole, D.H.; Sawyer, J.; Barlogie, B.; Tricot, G. Preceding standard therapy is the likely cause of MDS after autotransplants for multiple myeloma. Br. J. Haematol. 1996, 95, 349–353. [Google Scholar] [CrossRef]
  69. Kyle, R.A.; Pierre, R.V.; Bayrd, E.D. Multiple Myeloma and Acute Myelomonocytic Leukemia. N. Engl. J. Med. 1970, 283, 1121–1125. [Google Scholar] [CrossRef]
  70. Pemmaraju, N.; Shah, D.; Kantarjian, H.; Orlowski, R.Z.; Nogueras González, G.M.; Baladandayuthapani, V.; Jain, N.; Wagner, V.; Garcia-Manero, G.; Shah, J.; et al. Characteristics and outcomes of patients with multiple myeloma who develop therapy-related myelodysplastic syndrome, chronic myelomonocytic leukemia, or acute myeloid leukemia. Clin. Lymphoma Myeloma Leuk. 2015, 15, 110–114. [Google Scholar] [CrossRef]
  71. Palumbo, A.; Bringhen, S.; Kumar, S.K.; Lupparelli, G.; Usmani, S.; Waage, A.; Larocca, A.; van der Holt, B.; Musto, P.; Offidani, M.; et al. Second primary malignancies with lenalidomide therapy for newly diagnosed myeloma: A meta-analysis of individual patient data. Lancet Oncol. 2014, 15, 333–342. [Google Scholar] [CrossRef]
  72. Finnish Leukaemia Group. Acute leukaemia and other secondary neoplasms in patients treated with conventional chemotherapy for multiple myeloma: A Finnish Leukaemia Group study. Eur. J. Haematol. 2000, 65, 123–127. [Google Scholar] [CrossRef]
  73. Usmani, S.Z.; Sawyer, J.; Rosenthal, A.; Cottler-Fox, M.; Epstein, J.; Yaccoby, S.; Sexton, R.; Hoering, A.; Singh, Z.; Heuck, C.J.; et al. Risk factors for MDS and acute leukemia following total therapy 2 and 3 for multiple myeloma. Blood 2013, 121, 4753–4757. [Google Scholar] [CrossRef]
  74. Dimopoulos, M.A.; Richardson, P.G.; Brandenburg, N.; Yu, Z.; Weber, D.M.; Niesvizky, R.; Morgan, G.J. A review of second primary malignancy in patients with relapsed or refractory multiple myeloma treated with lenalidomide. Blood 2012, 119, 2764–2767. [Google Scholar] [CrossRef]
  75. Mailankody, S.; Pfeiffer, R.M.; Kristinsson, S.Y.; Korde, N.; Bjorkholm, M.; Goldin, L.R.; Turesson, I.; Landgren, O. Risk of acute myeloid leukemia and myelodysplastic syndromes after multiple myeloma and its precursor disease (MGUS). Blood 2011, 118, 4086–4092. [Google Scholar] [CrossRef]
  76. Roeker, L.E.; Larson, D.R.; Kyle, R.A.; Kumar, S.; Dispenzieri, A.; Rajkumar, S.V. Risk of acute leukemia and myelodysplastic syndromes in patients with monoclonal gammopathy of undetermined significance (MGUS): A population-based study of 17 315 patients. Leukemia 2013, 27, 1391–1393. [Google Scholar] [CrossRef]
  77. Turesson, I.; Kovalchik, S.A.; Pfeiffer, R.M.; Kristinsson, S.Y.; Goldin, L.R.; Drayson, M.T.; Landgren, O. Monoclonal gammopathy of undetermined significance and risk of lymphoid and myeloid malignancies: 728 cases followed up to 30 years in Sweden. Blood 2014, 123, 338–345. [Google Scholar] [CrossRef] [PubMed]
  78. Manier, S.; Salem, K.Z.; Park, J.; Landau, D.A.; Getz, G.; Ghobrial, I.M. Genomic complexity of multiple myeloma and its clinical implications. Nat. Rev. Clin. Oncol. 2017, 14, 100–113. [Google Scholar] [CrossRef]
  79. Maura, F.; Rustad, E.H.; Boyle, E.M.; Morgan, G.J. Reconstructing the evolutionary history of multiple myeloma. Best Pract. Res. Clin. Haematol. 2020, 33, 101145. [Google Scholar] [CrossRef] [PubMed]
  80. Bolli, N.; Genuardi, E.; Ziccheddu, B.; Martello, M.; Oliva, S.; Terragna, C. Next-Generation Sequencing for Clinical Management of Multiple Myeloma: Ready for Prime Time? Front. Oncol. 2020, 10, 189. [Google Scholar] [CrossRef]
  81. Chapman, M.A.; Lawrence, M.S.; Keats, J.J.; Cibulskis, K.; Sougnez, C.; Schinzel, A.C.; Harview, C.L.; Brunet, J.P.; Ahmann, G.J.; Adli, M.; et al. Initial genome sequencing and analysis of multiple myeloma. Nature 2011, 471, 467–472. [Google Scholar] [CrossRef]
  82. Walker, B.A.; Boyle, E.M.; Wardell, C.P.; Murison, A.; Begum, D.B.; Dahir, N.M.; Proszek, P.Z.; Johnson, D.C.; Kaiser, M.F.; Melchor, L.; et al. Mutational Spectrum, Copy Number Changes, and Outcome: Results of a Sequencing Study of Patients With Newly Diagnosed Myeloma. J. Clin. Oncol. 2015, 33, 3911–3920. [Google Scholar] [CrossRef]
  83. Rustad, E.H.; Yellapantula, V.; Glodzik, D.; Gundem, G.; Leongamornlert, D.A.; Campbell, P.J.; Papaemmanuil, E.; Landgren, O.; Maura, F. Revealing the Impact of Recurrent and Rare Structural Variations in Multiple Myeloma. Blood 2019, 134, 576. [Google Scholar] [CrossRef]
  84. Keats, J.J.; Craig, D.W.; Liang, W.; Venkata, Y.; Kurdoglu, A.; Aldrich, J.; Auclair, D.; Allen, K.; Harrison, B.; Jewell, S.; et al. Interim Analysis Of The MMRF CoMMpass Trial, a Longitudinal Study In Multiple Myeloma Relating Clinical Outcomes To Genomic and Immunophenotypic Profiles. Blood 2013, 122, 532. [Google Scholar] [CrossRef]
  85. Walker, B.A.; Samur, K.M.; Mavrommatis, K.; Ashby, C.; Wardell, P.C.; Ortiz, M.; Towfic, F.; Stein, K.C.; Bauer, A.M. The Multiple Myeloma Genome Project: Development of a Molecular Segmentation Strategy for the Clinical Classification of Multiple Myeloma. Blood 2016, 128, 196. [Google Scholar] [CrossRef]
  86. Walker, B.A.; Mavrommatis, K.; Wardell, C.P.; Ashby, T.C.; Bauer, M.; Davies, F.E.; Rosenthal, A.; Wang, H.; Qu, P.; Hoering, A.; et al. Identification of novel mutational drivers reveals oncogene dependencies in multiple myeloma. Blood 2018, 132, 587–597. [Google Scholar] [CrossRef]
  87. Rajkumar, S.V.; Kumar, S. Multiple myeloma current treatment algorithms. Blood Cancer J. 2020, 10, 94. [Google Scholar] [CrossRef] [PubMed]
  88. Dimopoulos, M.A.; Moreau, P.; Terpos, E.; Mateos, M.-V.; Zweegman, S.; Cook, G.; Delforge, M.; Hájek, R.; Schjesvold, F.; Cavo, M.; et al. Multiple Myeloma: EHA-ESMO Clinical Practice Guidelines for Diagnosis, Treatment and Follow-up. HemaSphere 2021, 5, e528. [Google Scholar] [CrossRef] [PubMed]
  89. Mohty, M.; Knauf, W.; Romanus, D.; Corman, S.; Verleger, K.; Kwon, Y.; Cherepanov, D.; Cambron-Mellott, M.J.; Vikis, H.G.; Gonzalez, F.; et al. Real-world treatment patterns and outcomes in non-transplant newly diagnosed multiple Myeloma in France, Germany, Italy, and the United Kingdom. Eur. J. Haematol. 2020, 105, 308–325. [Google Scholar] [CrossRef]
  90. Mateos, M.V.; Ludwig, H.; Bazarbachi, A.; Beksac, M.; Bladé, J.; Boccadoro, M.; Cavo, M.; Delforge, M.; Dimopoulos, M.A.; Facon, T.; et al. Insights on Multiple Myeloma Treatment Strategies. HemaSphere 2019, 3, e163. [Google Scholar] [CrossRef]
  91. Bobin, A.; Liuu, E.; Moya, N.; Gruchet, C.; Sabirou, F.; Lévy, A.; Gardeney, H.; Nsiala, L.; Cailly, L.; Guidez, S.; et al. Multiple myeloma: An overview of the current and novel therapeutic approaches in 2020. Cancers 2020, 12, 2885. [Google Scholar] [CrossRef]
  92. Muchtar, E.; Dingli, D.; Kumar, S.; Buadi, F.K.; Dispenzieri, A.; Hayman, S.R.; Wolf, R.C.; Gastineau, D.A.; Chakraborty, R.; Hogan, W.J.; et al. Autologous stem cell transplant for multiple myeloma patients 70 years or older. Bone Marrow Transplant. 2016, 51, 1449–1455. [Google Scholar] [CrossRef]
  93. Fiala, M.A.; King, J.; Feinberg, D.; Goldsmith, S.R.; Schroeder, M.A.; Ghobadi, A.; Stockerl-Goldstein, K.E.; Vij, R.; Wildes, T.M. Autologous stem cell transplant for patients with multiple myeloma between ages 75 and 78. Bone Marrow Transplant. 2021, 56, 2016–2018. [Google Scholar] [CrossRef] [PubMed]
  94. Vaxman, I.; Visram, A.; Kumar, S.; Dispenzieri, A.; Buadi, F.; Dingli, D.; Lacy, M.; Muchtar, E.; Kapoor, P.; Hogan, W.; et al. Autologous stem cell transplantation for multiple myeloma patients aged ≥ 75 treated with novel agents. Bone Marrow Transplant. 2021, 56, 1144–1150. [Google Scholar] [CrossRef] [PubMed]
  95. Dong, N.; McKiernan, P.; Siegel, D.S.D.; Vesole, D.H.; Rowley, S.D.; Andrews, T.; Ortega, A.; Skarbnik, A.P.; Biran, N.; Richter, J.R.; et al. Autologous stem cell transplantation in multiple myeloma patients over age 75. J. Clin. Oncol. 2018, 36, 8025. [Google Scholar] [CrossRef]
  96. Josepht, N.S.; Gupta, V.A.; Wyman, S.; Graiser, M.; Kaufman, J.L.; Almaula, D.; Andrews, J.; Hofmeister, C.; Dhodapkar, M.; Heffner, L.T.; et al. Benefits of autologous stem cell transplant for elderly myeloma patients in the last quarter of life. Transplant. Cell. Ther. 2021, in press. [Google Scholar]
  97. Palumbo, A.; Cavallo, F.; Gay, F.; Di Raimondo, F.; Ben Yehuda, D.; Petrucci, M.T.; Pezzatti, S.; Caravita, T.; Cerrato, C.; Ribakovsky, E.; et al. Autologous Transplantation and Maintenance Therapy in Multiple Myeloma. N. Engl. J. Med. 2014, 371, 895–905. [Google Scholar] [CrossRef] [PubMed]
  98. Attal, M.; Harousseau, J.-L.; Stoppa, A.-M.; Sotto, J.-J.; Fuzibet, J.-G.; Rossi, J.-F.; Casassus, P.; Maisonneuve, H.; Facon, T.; Ifrah, N.; et al. A Prospective, Randomized Trial of Autologous Bone Marrow Transplantation and Chemotherapy in Multiple Myeloma. N. Engl. J. Med. 1996, 335, 91–97. [Google Scholar] [CrossRef]
  99. Ntanasis-Stathopoulos, I.; Gavriatopoulou, M.; Kastritis, E.; Terpos, E.; Dimopoulos, M.A. Multiple myeloma: Role of autologous transplantation. Cancer Treat. Rev. 2020, 82, 101929. [Google Scholar] [CrossRef]
  100. Cavo, M.; Gay, F.; Beksac, M.; Pantani, L.; Petrucci, M.T.; Dimopoulos, M.A.; Dozza, L.; van der Holt, B.; Zweegman, S.; Oliva, S.; et al. Autologous haematopoietic stem-cell transplantation versus bortezomib–melphalan–prednisone, with or without bortezomib–lenalidomide–dexamethasone consolidation therapy, and lenalidomide maintenance for newly diagnosed multiple myeloma (EMN02/HO95): A mult. Lancet Haematol. 2020, 7, e456–e468. [Google Scholar] [CrossRef]
  101. Stadtmauer, E.A.; Pasquini, M.C.; Blackwell, B.; Hari, P.; Bashey, A.; Devine, S.; Efebera, Y.; Ganguly, S.; Gasparetto, C.; Geller, N.; et al. Autologous transplantation, consolidation, and maintenance therapy in multiple myeloma: Results of the BMT CTN 0702 trial. J. Clin. Oncol. 2019, 37, 589–597. [Google Scholar] [CrossRef]
  102. Gonsalves, W.I.; Buadi, F.K.; Ailawadhi, S.; Bergsagel, P.L.; Chanan Khan, A.A.; Dingli, D.; Dispenzieri, A.; Fonseca, R.; Hayman, S.R.; Kapoor, P.; et al. Utilization of hematopoietic stem cell transplantation for the treatment of multiple myeloma: A Mayo Stratification of Myeloma and Risk-Adapted Therapy (mSMART) consensus statement. Bone Marrow Transplant. 2019, 54, 353–367. [Google Scholar] [CrossRef]
  103. Legarda, M.A.; Cejalvo, M.J.; de la Rubia, J. Recent advances in the treatment of patients with multiple myeloma. Cancers 2020, 12, 3576. [Google Scholar] [CrossRef] [PubMed]
  104. Kumar, S.K.; Buadi, F.K.; Vincent Rajkumar, S. Pros and cons of frontline autologous transplant in multiple myeloma: The debate over timing. Blood 2019, 133, 652–659. [Google Scholar] [CrossRef] [PubMed]
  105. Greenberg, A.J.; Rajkumar, S.V.; Vachon, C.M. Familial monoclonal gammopathy of undetermined significance and multiple myeloma: Epidemiology, risk factors, and biological characteristics. Blood 2012, 119, 5359–5366. [Google Scholar] [CrossRef]
  106. Alexanian, R.; Haut, A.; Khan, A.U.; Lane, M.; McKelvey, E.M.; Migliore, P.J.; Stuckey, W.J., Jr.; Wilson, H.E. Treatment for Multiple Myeloma: Combination Chemotherapy With Different Melphalan Dose Regimens. JAMA 1969, 208, 1680–1685. [Google Scholar] [CrossRef]
  107. Bensinger, W.I.; Becker, P.S.; Gooley, T.A.; Chauncey, T.R.; Maloney, D.G.; Gopal, A.K.; Green, D.J.; Press, O.W.; Lill, M.; Ifthikharuddin, J.J.; et al. A randomized study of melphalan 200 mg/m2 vs 280 mg/m2 as a preparative regimen for patients with multiple myeloma undergoing auto-SCT. Bone Marrow Transplant. 2016, 51, 67–71. [Google Scholar] [CrossRef]
  108. Palumbo, A.; Bringhen, S.; Bruno, B.; Falcone, A.P.; Liberati, A.M.; Grasso, M.; Ria, R.; Pisani, F.; Cangialosi, C.; Caravita, T.; et al. Melphalan 200 mg/m2 versus melphalan 100 mg/m2 in newly diagnosed myeloma patients: A prospective, multicenter phase 3 study. Blood 2010, 115, 1873–1879. [Google Scholar] [CrossRef]
  109. Gay, F.; Oliva, S.; Petrucci, M.T.; Conticello, C.; Catalano, L.; Corradini, P.; Siniscalchi, A.; Magarotto, V.; Pour, L.; Carella, A.; et al. Chemotherapy plus lenalidomide versus autologous transplantation, followed by lenalidomide plus prednisone versus lenalidomide maintenance, in patients with multiple myeloma: A randomised, multicentre, phase 3 trial. Lancet Oncol. 2015, 16, 1617–1629. [Google Scholar] [CrossRef]
  110. Kazandjian, D.; Mo, C.C.; Landgren, O.; Richardson, P.G. The role of high-dose melphalan with autologous stem-cell transplant in multiple myeloma: Is it time for a paradigm shift? Br. J. Haematol. 2020, 191, 692–703. [Google Scholar] [CrossRef]
  111. Derudas, D.; Capraro, F.; Martinelli, G.; Cerchione, C. How i manage frontline transplant-ineligible multiple myeloma. Hematol. Rep. 2020, 12 (Suppl. 1), 8956. [Google Scholar] [CrossRef]
  112. Facon, T.; Kumar, S.; Plesner, T.; Orlowski, R.Z.; Moreau, P.; Bahlis, N.; Basu, S.; Nahi, H.; Hulin, C.; Quach, H.; et al. Daratumumab plus Lenalidomide and Dexamethasone for Untreated Myeloma. N. Engl. J. Med. 2019, 380, 2104–2115. [Google Scholar] [CrossRef]
  113. Dimopoulos, M.A.; Jakubowiak, A.J.; McCarthy, P.L.; Orlowski, R.Z.; Attal, M.; Bladé, J.; Goldschmidt, H.; Weisel, K.C.; Ramasamy, K.; Zweegman, S.; et al. Developments in continuous therapy and maintenance treatment approaches for patients with newly diagnosed multiple myeloma. Blood Cancer J. 2020, 10, 1–19. [Google Scholar] [CrossRef]
  114. Cook, G.; Ashcroft, A.J.; Cairns, D.A.; Williams, C.D.; Brown, J.M.; Cavenagh, J.D.; Snowden, J.A.; Parrish, C.; Yong, K.; Cavet, J.; et al. The effect of salvage autologous stem-cell transplantation on overall survival in patients with relapsed multiple myeloma (final results from BSBMT/UKMF Myeloma X Relapse [Intensive]): A randomised, open-label, phase 3 trial. Lancet Haematol. 2016, 3, e340–e351. [Google Scholar] [CrossRef]
  115. Cavo, M.; Tosi, P.; Zamagni, E.; Cellini, C.; Tacchetti, P.; Patriarca, F.; Di Raimondo, F.; Volpe, E.; Ronconi, S.; Cangini, D.; et al. Prospective, randomized study of single compared with double autologous stem-cell transplantation for multiple myeloma: Bologna 96 clinical study. J. Clin. Oncol. 2007, 25, 2434–2441. [Google Scholar] [CrossRef]
  116. Kropff, M.; Vogel, M.; Bisping, G.; Schlag, R.; Weide, R.; Knauf, W.; Fiechtner, H.; Kojouharoff, G.; Kremers, S.; Berdel, W.E. Bortezomib and low-dose dexamethasone with or without continuous low-dose oral cyclophosphamide for primary refractory or relapsed multiple myeloma: A randomized phase III study. Ann. Hematol. 2017, 96, 1857–1866. [Google Scholar] [CrossRef]
  117. Reeder, C.B.; Reece, D.E.; Kukreti, V.; Mikhael, J.R.; Chen, C.; Trudel, S.; Laumann, K.; Vohra, H.; Fonseca, R.; Bergsagel, P.L.; et al. Long-term survival with cyclophosphamide, bortezomib and dexamethasone induction therapy in patients with newly diagnosed multiple myeloma. Br. J. Haematol. 2014, 167, 563–565. [Google Scholar] [CrossRef]
  118. Garderet, L.; Kuhnowski, F.; Berge, B.; Roussel, M.; Escoffre-Barbe, M.; Lafon, I.; Facon, T.; Leleu, X.; Karlin, L.; Perrot, A.; et al. Pomalidomide, cyclophosphamide, and dexamethasone for relapsed multiple myeloma. Blood 2018, 132, 2555–2563. [Google Scholar] [CrossRef] [PubMed]
  119. Morgan, G.J.; Davies, F.E.; Gregory, W.M.; Russell, N.H.; Bell, S.E.; Szubert, A.J.; Coy, N.N.; Cook, G.; Feyler, S.; Byrne, J.L.; et al. Cyclophosphamide, thalidomide, and dexamethasone (CTD) as initial therapy for patients with multiple myeloma unsuitable for autologous transplantation. Blood 2011, 118, 1231–1238. [Google Scholar] [CrossRef] [PubMed]
  120. Kumar, S.; Flinn, I.; Richardson, P.G.; Hari, P.; Callander, N.; Noga, S.J.; Stewart, A.K.; Turturro, F.; Rifkin, R.; Wolf, J.; et al. Randomized, multicenter, phase 2 study (EVOLUTION) of combinations of bortezomib, dexamethasone, cyclophosphamide, and lenalidomide in previously untreated multiple myeloma. Blood 2012, 119, 4375–4382. [Google Scholar] [CrossRef] [PubMed]
  121. Yimer, H.; Melear, J.; Faber, E.; Bensinger, W.I.; Burke, J.M.; Narang, M.; Stevens, D.; Gunawardena, S.; Lutska, Y.; Qi, K.; et al. Daratumumab, bortezomib, cyclophosphamide and dexamethasone in newly diagnosed and relapsed multiple myeloma: LYRA study. Br. J. Haematol. 2019, 185, 492–502. [Google Scholar] [CrossRef]
  122. Jimenez Zepeda, V.H.; Duggan, P.; Neri, P.E.; Bahlis, N.J. Cyclophosphamide, Bortezomib and Dexamethasone (CyBORD) Is a Feasible and Active Regimen for Non-Transplant Eligible Multiple Myeloma Patients. Blood 2014, 124, 5751. [Google Scholar] [CrossRef]
  123. Reeder, C.B.; Reece, D.E.; Kukreti, V.; Chen, C.; Trudel, S.; Hentz, J.; Noble, B.; Pirooz, N.A.; Spong, J.E.; Piza, J.G.; et al. Cyclophosphamide, bortezomib and dexamethasone induction for newly diagnosed multiple myeloma: High response rates in a phase II clinical trial. Leukemia 2009, 23, 1337–1341. [Google Scholar] [CrossRef]
  124. Areethamsirikul, N.; Masih-Khan, E.; Chu, C.M.; Jimenez-Zepeda, V.; Reece, D.E.; Trudel, S.; Kukreti, V.; Tiedemann, R.; Chen, C. CyBorD induction therapy in clinical practice. Bone Marrow Transplant. 2015, 50, 375–379. [Google Scholar] [CrossRef] [PubMed]
  125. Khan, M.L.; Reeder, C.B.; Kumar, S.; Lacy, M.; Reece, D.E.; Laumann, K.M.; Mikhael, J.; Fonseca, R.; Rajkumar, V.; Stewart, A.K. A comparison of lenalidomide/dexamethasone (RD) versus cyclophosphamide/lenalidomide/dexamethasone (CRD) versus cyclophosphamide/bortezomib/dexamethasone (CyborD) in newly diagnosed multiple myeloma (MM). J. Clin. Oncol. 2010, 28, 8131. [Google Scholar] [CrossRef]
  126. Pinto, V.; Bergantim, R.; Caires, H.R.; Seca, H.; Guimarães, J.E.; Vasconcelos, M.H. Multiple myeloma: Available therapies and causes of drug resistance. Cancers 2020, 12, 407. [Google Scholar] [CrossRef] [PubMed]
  127. Chari, A.; Vogl, D.T.; Gavriatopoulou, M.; Nooka, A.K.; Yee, A.J.; Huff, C.A.; Moreau, P.; Dingli, D.; Cole, C.; Lonial, S.; et al. Oral Selinexor–Dexamethasone for Triple-Class Refractory Multiple Myeloma. N. Engl. J. Med. 2019, 381, 727–738. [Google Scholar] [CrossRef] [PubMed]
  128. Lonial, S.; Lee, H.C.; Badros, A.; Trudel, S.; Nooka, A.K.; Chari, A.; Abdallah, A.O.; Callander, N.; Lendvai, N.; Sborov, D.; et al. Belantamab mafodotin for relapsed or refractory multiple myeloma (DREAMM-2): A two-arm, randomised, open-label, phase 2 study. Lancet Oncol. 2020, 21, 207–221. [Google Scholar] [CrossRef]
  129. Shah, J.J.; Kaufman, J.L.; Zonder, J.A.; Cohen, A.D.; Bensinger, W.I.; Hilder, B.W.; Rush, S.A.; Walker, D.H.; Tunquist, B.J.; Litwiler, K.S.; et al. A Phase 1 and 2 study of Filanesib alone and in combination with low-dose dexamethasone in relapsed/refractory multiple myeloma. Cancer 2017, 123, 4617–4630. [Google Scholar] [CrossRef]
  130. Mikkilineni, L.; Kochenderfer, J.N. CAR T cell therapies for patients with multiple myeloma. Nat. Rev. Clin. Oncol. 2021, 18, 71–84. [Google Scholar] [CrossRef]
  131. van de Donk, N.W.C.J.; Usmani, S.Z.; Yong, K. CAR T-cell therapy for multiple myeloma: State of the art and prospects. Lancet Haematol. 2021, 8, e446–e461. [Google Scholar] [CrossRef]
  132. Offidani, M.; Corvatta, L.; Morè, S.; Olivieri, A. Novel experimental drugs for treatment of multiple myeloma. J. Exp. Pharmacol. 2021, 13, 245–264. [Google Scholar] [CrossRef] [PubMed]
  133. Lassiter, G.; Bergeron, C.; Guedry, R.; Cucarola, J.; Kaye, A.M.; Cornett, E.M.; Kaye, A.D.; Varrassi, G.; Viswanath, O.; Urits, I. Belantamab mafodotin to treat multiple myeloma: A comprehensive review of disease, drug efficacy and side effects. Curr. Oncol. 2021, 28, 63. [Google Scholar] [CrossRef]
  134. Raje, N.; Berdeja, J.; Lin, Y.; Siegel, D.; Jagannath, S.; Madduri, D.; Liedtke, M.; Rosenblatt, J.; Maus, M.V.; Turka, A.; et al. Anti-BCMA CAR T-Cell Therapy bb2121 in Relapsed or Refractory Multiple Myeloma. N. Engl. J. Med. 2019, 380, 1726–1737. [Google Scholar] [CrossRef] [PubMed]
  135. Gourzones, C.; Bret, C.; Moreaux, J. Treatment May Be Harmful: Mechanisms/Prediction/Prevention of Drug-Induced DNA Damage and Repair in Multiple Myeloma. Front. Genet. 2019, 10, 861. [Google Scholar] [CrossRef]
  136. Esma, F.; Salvini, M.; Troia, R.; Boccadoro, M.; Larocca, A.; Pautasso, C. Melphalan hydrochloride for the treatment of multiple myeloma. Expert Opin. Pharmacother. 2017, 18, 1127–1136. [Google Scholar] [CrossRef]
  137. Osborne, M.R.; Lawley, P.D. Alkylation of DNA by melphalan with special reference to adenine derivatives and adenine-guanine cross-linking. Chem. Biol. Interact. 1993, 89, 49–60. [Google Scholar] [CrossRef]
  138. Osborne, M.R.; Lawley, P.D. The reaction of melphalan with deoxyguanosine and deoxyguanylic acid. Chem. Biol. Interact. 1992, 84, 189–198. [Google Scholar] [CrossRef]
  139. Musto, P.; D’Auria, F. Melphalan: Old and new uses of a still master drug for multiple myeloma. Expert Opin. Investig. Drugs 2007, 16, 1467–1487. [Google Scholar] [CrossRef]
  140. Schjesvold, F.; Oriol, A. Current and novel alkylators in multiple myeloma. Cancers 2021, 13, 2465. [Google Scholar] [CrossRef]
  141. Mehta, J.R.; Przybylski, M.; Ludlum, D.B. Alkylation of Guanosine and Deoxyguanosine by Phosphoramide Mustard. Cancer Res. 1980, 40, 4183–4186. [Google Scholar]
  142. Maccubbin, A.E.; C’aballes, L.; Gurtoo, H.L. A Cyclophosphamide/DNA Phosphoester Adduct Formed in Vitro and in Vivo. Cancer Res. 1991, 51, 886–892. [Google Scholar]
  143. Benson, A.J.; Martin, C.N.; Garner, R.C. N-(2-hydroxyethyl)-N-[2-(7-guaninyl)ethyl]amine, the putative major DNA adduct of cyclophosphamide in vitro and in vivo in the rat. Biochem. Pharmacol. 1988, 37, 2979–2985. [Google Scholar] [CrossRef]
  144. Swan, D.; Gurney, M.; Krawczyk, J.; Ryan, A.E.; O’Dwyer, M. Beyond DNA Damage: Exploring the Immunomodulatory Effects of Cyclophosphamide in Multiple Myeloma. HemaSphere 2020, 4, e350. [Google Scholar] [CrossRef]
  145. Landau, H.J.; Yellapantula, V.; Diamond, B.T.; Rustad, E.H.; Maclachlan, K.H.; Gundem, G.; Medina-Martinez, J.; Ossa, J.A.; Levine, M.F.; Zhou, Y.; et al. Accelerated single cell seeding in relapsed multiple myeloma. Nat. Commun. 2020, 11, 3617. [Google Scholar] [CrossRef] [PubMed]
  146. Maura, F.; Weinhold, N.; Diamond, B.; Kazandjian, D.; Rasche, L.; Morgan, G.; Landgren, O. The mutagenic impact of melphalan in multiple myeloma. Leukemia 2021, 35, 2145–2150. [Google Scholar] [CrossRef]
  147. Gonzalez, F.; Trujillo, J.M.; Alexanian, R. Acute leukemia in multiple myeloma. Ann. Intern. Med. 1977, 86, 440–443. [Google Scholar] [CrossRef] [PubMed]
  148. Law, I.P.; Blom, J. Second malignancies in patients with multiple myeloma. Oncology 1977, 34, 20–24. [Google Scholar] [CrossRef] [PubMed]
  149. Bergsagel, D.E.; Bailey, A.J.; Langley, G.R.; MacDonald, R.N.; White, D.F.; Miller, A.B. The Chemotherapy of Plasma-Cell Myeloma and the Incidence of Acute Leukemia. N. Engl. J. Med. 1979, 301, 743–748. [Google Scholar] [CrossRef] [PubMed]
  150. Rosner, F.; Grünwald, H. Multiple myeloma terminating in acute leukemia. Report of 12 cases and review of the literature. Am. J. Med. 1974, 57, 927–939. [Google Scholar] [CrossRef]
  151. Richardson, P.G.; Bringhen, S.; Voorhees, P.; Plesner, T.; Mellqvist, U.H.; Reeves, B.; Paba-Prada, C.; Zubair, H.; Byrne, C.; Chauhan, D.; et al. Melflufen plus dexamethasone in relapsed and refractory multiple myeloma (O-12-M1): A multicentre, international, open-label, phase 1–2 study. Lancet Haematol. 2020, 7, e395–e407. [Google Scholar] [CrossRef]
  152. Wickström, M.; Nygren, P.; Larsson, R.; Harmenberg, J.; Lindberg, J.; Sjöberg, P.; Jerling, M.; Lehmann, F.; Richardson, P.; Anderson, K.; et al. Melflufen-a peptidase-potentiated alkylating agent in clinical trials. Oncotarget 2017, 8, 66641–66655. [Google Scholar] [CrossRef]
  153. Mateos, M.V.; González-Calle, V. Is there a role for new drugs with alkylating properties in multiple myeloma? Lancet Haematol. 2020, 7, e357–e359. [Google Scholar] [CrossRef]
  154. Soll, J.M.; Sobol, R.W.; Mosammaparast, N. Regulation of DNA Alkylation Damage Repair: Lessons and Therapeutic Opportunities. Trends Biochem. Sci. 2017, 42, 206–218. [Google Scholar] [CrossRef]
  155. Friedberg, E.C.; Walker, G.C.; Siede, W.; Wood, R.D.; Schultz, R.A.; Ellenberger, T. DNA Repair and Mutagenesis, 2nd ed.; ASM Press: Washington, DC, USA, 2005. [Google Scholar]
  156. Datta, A.; Brosh, R.M. Holding all the cards—how fanconi anemia proteins deal with replication stress and preserve genomic stability. Genes 2019, 10, 170. [Google Scholar] [CrossRef] [PubMed]
  157. Ceccaldi, R.; Sarangi, P.; D’Andrea, A.D. The Fanconi anaemia pathway: New players and new functions. Nat. Rev. Mol. Cell Biol. 2016, 17, 337–349. [Google Scholar] [CrossRef] [PubMed]
  158. Renaudin, X.; Rosselli, F. The FANC/BRCA Pathway Releases Replication Blockades by Eliminating DNA Interstrand Cross-Links. Genes 2020, 11, 585. [Google Scholar] [CrossRef] [PubMed]
  159. Lyakhovich, A.; Surralles, J. Disruption of the Fanconi anemia/BRCA pathway in sporadic cancer. Cancer Lett. 2006, 232, 99–106. [Google Scholar] [CrossRef] [PubMed]
  160. Chen, Q.; Van Der Sluis, P.C.; Boulware, D.; Hazlehurst, L.A.; Dalton, W.S. The FA/BRCA pathway is involved in melphalan-induced DNA interstrand cross-link repair and accounts for melphalan resistance in multiple myeloma cells. Blood 2005, 106, 698–705. [Google Scholar] [CrossRef]
  161. Tessoulin, B.; Moreau-Aubry, A.; Descamps, G.; Gomez-Bougie, P.; Maïga, S.; Gaignard, A.; Chiron, D.; Ménoret, E.; Le Gouill, S.; Moreau, P.; et al. Whole-exon sequencing of human myeloma cell lines shows mutations related to myeloma patients at relapse with major hits in the DNA regulation and repair pathways. J. Hematol. Oncol. 2018, 11, 137. [Google Scholar] [CrossRef]
  162. Turner, J.G.; Cui, Y.; Bauer, A.A.; Dawson, J.L.; Gomez, J.A.; Kim, J.; Cubitt, C.L.; Nishihori, T.; Dalton, W.S.; Sullivan, D.M. Melphalan and Exportin 1 Inhibitors Exert Synergistic Antitumor Effects in Preclinical Models of Human Multiple Myeloma. Cancer Res. 2020, 80, 5344–5354. [Google Scholar] [CrossRef] [PubMed]
  163. Herrero, A.B.; SanMiguel, J.; Gutierrez, N.C. Deregulation of DNA double-strand break repair in multiple myeloma: Implications for genome stability. PLoS ONE 2015, 10, e0121581. [Google Scholar] [CrossRef]
  164. Herrero, A.B.; Gutiérrez, N.C. Targeting ongoing DNA damage in multiple myeloma: Effects of DNA damage response inhibitors on plasma cell survival. Front. Oncol. 2017, 7, 98. [Google Scholar] [CrossRef] [PubMed]
  165. Shammas, M.A.; Reis, R.J.S.; Koley, H.; Batchu, R.B.; Li, C.; Munshi, N.C. Dysfunctional homologous recombination mediates genomic instability and progression in myeloma. Blood 2009, 113, 2290–2297. [Google Scholar] [CrossRef]
  166. Spanswick, V.J.; Lowe, H.L.; Newton, C.; Bingham, J.P.; Bagnobianchi, A.; Kiakos, K.; Craddock, C.; Ledermann, J.A.; Hochhauser, D.; Hartley, J.A. Evidence for different mechanisms of “unhooking” for melphalan and cisplatin-induced DNA interstrand cross-links in vitro and in clinical acquired resistant tumour samples. BMC Cancer 2012, 12, 436. [Google Scholar] [CrossRef]
  167. Tageja, N.; Nagi, J. Bendamustine: Something old, something new. Cancer Chemother. Pharmacol. 2010, 66, 413–423. [Google Scholar] [CrossRef] [PubMed]
  168. Hiraoka, N.; Kikuchi, J.; Yamauchi, T.; Koyama, D.; Wada, T.; Uesawa, M.; Akutsu, M.; Mori, S.; Nakamura, Y.; Ueda, T.; et al. Purine analog-like properties of bendamustine underlie rapid activation of DNA damage response and synergistic effects with pyrimidine analogues in lymphoid malignancies. PLoS ONE 2014, 9, e90675. [Google Scholar]
  169. Cives, M.; Ciavarella, S.; Rizzo, F.M.; De Matteo, M.; Dammacco, F.; Silvestris, F. Bendamustine overcomes resistance to melphalan in myeloma cell lines by inducing cell death through mitotic catastrophe. Cell. Signal. 2013, 25, 1108–1117. [Google Scholar] [CrossRef]
  170. Gomez-Arteaga, A.; Mark, T.M.; Guarneri, D.; Christos, P.J.; Gergis, U.; Greenberg, J.D.; Hsu, J.; Mayer, S.A.; Niesvizky, R.; Pearse, R.N.; et al. High-dose bendamustine and melphalan conditioning for autologous stem cell transplantation for patients with multiple myeloma. Bone Marrow Transplant. 2019, 54, 2027–2038. [Google Scholar] [CrossRef]
  171. Gentile, M.; Vigna, E.; Recchia, A.G.; Morabito, L.; Mendicino, F.; Giagnuolo, G.; Morabito, F. Bendamustine in multiple myeloma. Eur. J. Haematol. 2015, 95, 377–388. [Google Scholar] [CrossRef]
  172. Mateos, M.V.; Orio, A.; Rosiño, L.; Arriba, F.D.; Puig, N.; Martín, J.; Martínez-López, J.; Echeveste, M.A.; Sarrá, J.; Ocio, E.; et al. Bendamustine, bortezomib and prednisone for the treatment of patients with newly diagnosed multiple myeloma: Results of a prospective phase 2 Spanish/PETHEMA trial. Haematologica 2015, 100, 1096–1102. [Google Scholar] [CrossRef]
  173. Rodon, P.; Hulin, C.; Pegourie, B.; Tiab, M.; Anglaret, B.; Benboubker, L.; Jardel, H.; Decaux, O.; Kolb, B.; Roussel, M.; et al. Phase ii study of bendamustine, bortezomib and dexamethasone as Second-Line treatment for elderly patients with multiple myeloma: The intergroupe francophone du myelome 2009-01 trial. Haematologica 2015, 100, e56–e59. [Google Scholar] [CrossRef]
  174. Knauf, W.; Dingeldein, G.; Schlag, R.; Welslau, M.; Moehler, T.; Terzer, T.; Walter, S.; Habermehl, C.; Kunz, C.; Goldschmidt, H.; et al. First-line therapy with bendamustine/prednisone/bortezomib—A GMMG trial for non-transplant eligible symptomatic multiple myeloma patients. Eur. J. Haematol. 2020, 105, 116–125. [Google Scholar] [CrossRef]
  175. Pönisch, W.; Mitrou, P.S.; Merkle, K.; Herold, M.; Assmann, M.; Wilhelm, G.; Dachselt, K.; Richter, P.; Schirmer, V.; Schulze, A.; et al. Treatment of Bendamustine and Prednisone in patients with newly diagnosed multiple myeloma results in superior complete response rate, prolonged time to treatment failure and improved quality of life compared to treatment with Melphalan and Prednisone—A randomized phase III study of the East German Study Group of Hematology and Oncology (OSHO). J. Cancer Res. Clin. Oncol. 2006, 132, 205–212. [Google Scholar] [PubMed]
  176. Mark, T.M.; Reid, W.; Niesvizky, R.; Gergis, U.; Pearse, R.; Mayer, S.; Greenberg, J.; Coleman, M.; Van Besien, K.; Shore, T. A phase 1 study of bendamustine and melphalan conditioning for autologous stem cell transplantation in multiple myeloma. Biol. Blood Marrow Transplant. 2013, 19, 831–837. [Google Scholar] [CrossRef] [PubMed]
  177. Asatsuma-Okumura, T.; Ito, T.; Handa, H. Molecular mechanisms of cereblon-based drugs. Pharmacol. Ther. 2019, 202, 132–139. [Google Scholar] [CrossRef]
  178. Moreau, P.; Hulin, C.; Macro, M.; Caillot, D.; Chaleteix, C.; Roussel, M.; Garderet, L.; Royer, B.; Brechignac, S.; Tiab, M.; et al. VTD is superior to VCD prior to intensive therapy in multiple myeloma: Results of the prospective IFM2013-04 trial. Blood 2016, 127, 2569–2574. [Google Scholar] [CrossRef]
  179. Richardson, P.G.; Weller, E.; Lonial, S.; Jakubowiak, A.J.; Jagannath, S.; Raje, N.S.; Avigan, D.E.; Xie, W.; Ghobrial, I.M.; Schlossman, R.L.; et al. Lenalidomide, bortezomib, and dexamethasone combination therapy in patients with newly diagnosed multiple myeloma. Blood 2010, 116, 679–686. [Google Scholar] [CrossRef] [PubMed]
  180. Miguel, J.S.; Weisel, K.; Moreau, P.; Lacy, M.; Song, K.; Delforge, M.; Karlin, L.; Goldschmidt, H.; Banos, A.; Oriol, A.; et al. Pomalidomide plus low-dose dexamethasone versus high-dose dexamethasone alone for patients with relapsed and refractory multiple myeloma (MM-003): A randomised, open-label, phase 3 trial. Lancet Oncol. 2013, 14, 1055–1066. [Google Scholar] [CrossRef]
  181. Siegel, D.S.; Schiller, G.J.; Song, K.W.; Agajanian, R.; Stockerl-Goldstein, K.; Kaya, H.; Sebag, M.; Samaras, C.; Malek, E.; Talamo, G.; et al. Pomalidomide plus low-dose dexamethasone in relapsed refractory multiple myeloma after lenalidomide treatment failure. Br. J. Haematol. 2020, 188, 501–510. [Google Scholar] [CrossRef]
  182. Terpos, E.; Repousis, P.; Lalayanni, C.; Hatjiharissi, E.; Assimakopoulou, T.; Vassilopoulos, G.; Pouli, A.; Spanoudakis, E.; Michalis, E.; Pangalis, G.; et al. Pomalidomide Plus Low-Dose Dexamethasone in Relapsed/Refractory Multiple Myeloma Patients: Results of the Real-World “POWERFUL” Study. J. Clin. Med. 2021, 10, 1509. [Google Scholar] [CrossRef]
  183. Barosi, G.; Gale, R.P. Is lenalidomide the standard-of-care after an autotransplant for plasma cell myeloma? Leukemia 2019, 33, 588–596. [Google Scholar] [CrossRef]
  184. Attal, M.; Lauwers-Cances, V.; Marit, G.; Caillot, D.; Moreau, P.; Facon, T.; Stoppa, A.M.; Hulin, C.; Benboubker, L.; Garderet, L.; et al. Lenalidomide Maintenance after Stem-Cell Transplantation for Multiple Myeloma. N. Engl. J. Med. 2012, 366, 1782–1791. [Google Scholar] [CrossRef] [PubMed]
  185. Jackson, G.H.; Davies, F.E.; Pawlyn, C.; Cairns, D.A.; Striha, A.; Collett, C.; Hockaday, A.; Jones, J.R.; Kishore, B.; Garg, M.; et al. Lenalidomide maintenance versus observation for patients with newly diagnosed multiple myeloma (Myeloma XI): A multicentre, open-label, randomised, phase 3 trial. Lancet Oncol. 2019, 20, 57–73. [Google Scholar] [CrossRef]
  186. McCarthy, P.L.; Holstein, S.A.; Petrucci, M.T.; Richardson, P.G.; Hulin, C.; Tosi, P.; Bringhen, S.; Musto, P.; Anderson, K.C.; Caillot, D.; et al. Lenalidomide maintenance after autologous stem-cell transplantation in newly diagnosed multiple myeloma: A meta-analysis. J. Clin. Oncol. 2017, 35, 3279–3289. [Google Scholar] [CrossRef]
  187. Palumbo, A.; Hajek, R.; Delforge, M.; Kropff, M.; Petrucci, M.T.; Catalano, J.; Gisslinger, H.; Wiktor-Jędrzejczak, W.; Zodelava, M.; Weisel, K.; et al. Continuous Lenalidomide Treatment for Newly Diagnosed Multiple Myeloma. N. Engl. J. Med. 2012, 366, 1759–1769. [Google Scholar] [CrossRef] [PubMed]
  188. McCarthy, P.L.; Owzar, K.; Hofmeister, C.C.; Hurd, D.D.; Hassoun, H.; Richardson, P.G.; Giralt, S.; Stadtmauer, E.A.; Weisdorf, D.J.; Vij, R.; et al. Lenalidomide after Stem-Cell Transplantation for Multiple Myeloma. N. Engl. J. Med. 2012, 366, 1770–1781. [Google Scholar] [CrossRef]
  189. Ito, T.; Ando, H.; Suzuki, T.; Ogura, T.; Hotta, K.; Imamura, Y.; Yamaguchi, Y.; Handa, H. Identification of a primary target of thalidomide teratogenicity. Science 2010, 327, 1345–1350. [Google Scholar] [CrossRef] [PubMed]
  190. Lopez-Girona, A.; Mendy, D.; Ito, T.; Miller, K.; Gandhi, A.K.; Kang, J.; Karasawa, S.; Carmel, G.; Jackson, P.; Abbasian, M.; et al. Cereblon is a direct protein target for immunomodulatory and antiproliferative activities of lenalidomide and pomalidomide. Leukemia 2012, 26, 2326–2335. [Google Scholar] [CrossRef]
  191. Fischer, E.S.; Böhm, K.; Lydeard, J.R.; Yang, H.; Stadler, M.B.; Cavadini, S.; Nagel, J.; Serluca, F.; Acker, V.; Lingaraju, G.M.; et al. Structure of the DDB1-CRBN E3 ubiquitin ligase in complex with thalidomide. Nature 2014, 512, 49–53. [Google Scholar] [CrossRef]
  192. Zhu, Y.X.; Braggio, E.; Shi, C.X.; Bruins, L.A.; Schmidt, J.E.; Van Wier, S.; Chang, X.B.; Bjorklund, C.C.; Fonseca, R.; Bergsagel, P.L.; et al. Cereblon expression is required for the antimyeloma activity of lenalidomide and pomalidomide. Blood 2011, 118, 4771–4779. [Google Scholar] [CrossRef]
  193. Abruzzese, M.P.; Bilotta, M.T.; Fionda, C.; Zingoni, A.; Soriani, A.; Petrucci, M.T.; Ricciardi, M.R.; Molfetta, R.; Paolini, R.; Santoni, A.; et al. The homeobox transcription factor MEIS2 is a regulator of cancer cell survival and IMiDs activity in Multiple Myeloma: Modulation by Bromodomain and Extra-Terminal (BET) protein inhibitors. Cell Death Dis. 2019, 10, 324. [Google Scholar] [CrossRef]
  194. Krönke, J.; Udeshi, N.D.; Narla, A.; Grauman, P.; Hurst, S.N.; McConkey, M.; Svinkina, T.; Heckl, D.; Comer, E.; Li, X.; et al. Lenalidomide causes selective degradation of IKZF1 and IKZF3 in multiple myeloma cells. Science 2014, 343, 301–305. [Google Scholar] [CrossRef] [PubMed]
  195. Lu, G.; Middleton, R.E.; Sun, H.; Naniong, M.V.; Ott, C.J.; Mitsiades, C.S.; Wong, K.K.; Bradner, J.E.; Kaelin, W.G. The myeloma drug lenalidomide promotes the cereblon-dependent destruction of ikaros proteins. Science 2014, 343, 305–309. [Google Scholar] [CrossRef]
  196. Zhu, Y.X.; Braggio, E.; Shi, C.X.; Kortuem, K.M.; Bruins, L.A.; Schmidt, J.E.; Chang, X.B.; Langlais, P.; Luo, M.; Jedlowski, P.; et al. Identification of cereblon-binding proteins and relationship with response and survival after IMiDs in multiple myeloma. Blood 2014, 124, 536–545. [Google Scholar] [CrossRef] [PubMed]
  197. Agnarelli, A.; Chevassut, T.; Mancini, E.J. IRF4 in multiple myeloma—Biology, disease and therapeutic target. Leuk. Res. 2018, 72, 52–58. [Google Scholar] [CrossRef]
  198. Jovanović, K.K.; Roche-Lestienne, C.; Ghobrial, I.M.; Facon, T.; Quesnel, B.; Manier, S. Targeting MYC in multiple myeloma. Leukemia 2018, 32, 1295–1306. [Google Scholar] [CrossRef] [PubMed]
  199. Shaffer, A.L.; Tolga Emre, N.C.; Romesser, P.B.; Staudt, L.M. IRF4: Immunity. Malignancy! Therapy? Clin. Cancer Res. 2009, 15, 2954–2961. [Google Scholar] [CrossRef]
  200. Klein, U.; Casola, S.; Cattoretti, G.; Shen, Q.; Lia, M.; Mo, T.; Ludwig, T.; Rajewsky, K.; Dalla-Favera, R. Transcription factor IRF4 controls plasma cell differentiation and class-switch recombination. Nat. Immunol. 2006, 7, 773–782. [Google Scholar] [CrossRef]
  201. Iida, S.; Rao, P.H.; Butler, M.; Corradini, P.; Boccadoro, M.; Klein, B.; Chaganti, R.S.K.; Dalla-Favera, R. Deregulation of MUM1/IRF4 by chromosomal translocation in multiple myeloma. Nat. Genet. 1997, 17, 230. [Google Scholar] [CrossRef]
  202. Yoshida, S.; Nakazawa, N.; Iida, S.; Hayami, Y.; Sato, S.; Wakita, A.; Shimizu, S.; Taniwaki, M.; Ueda, R. Detection of MUM1/IRF4-IgH fusion in multiple myeloma. Leukemia 1999, 13, 1812–1816. [Google Scholar] [CrossRef]
  203. Wang, L.; Yao, Z.Q.; Moorman, J.P.; Xu, Y.; Ning, S. Gene expression profiling identifies IRF4-Associated molecular signatures in hematological malignancies. PLoS ONE 2014, 9, e106788. [Google Scholar] [CrossRef] [PubMed]
  204. Misiewicz-Krzeminska, I.; de Ramón, C.; Corchete, L.A.; Krzeminski, P.; Rojas, E.A.; Isidro, I.; García-Sanz, R.; Martínez-López, J.; Oriol, A.; Bladé, J.; et al. Quantitative expression of Ikaros, IRF4, and PSMD10 proteins predicts survival in VRD-treated patients with multiple myeloma. Blood Adv. 2020, 4, 6023–6033. [Google Scholar] [CrossRef]
  205. Moreira, A.L.; Sampaio, E.P.; Zmuidzinas, A.; Frindt, P.; Smith, K.A.; Kaplan, G. Thalidomide exerts its inhibitory action on tumor necrosis factor α by enhancing mrna degradation. J. Exp. Med. 1993, 177, 1675–1680. [Google Scholar] [CrossRef] [PubMed]
  206. Payvandi, F.; Wu, L.; Haley, M.; Schafer, P.H.; Zhang, L.H.; Chen, R.S.; Muller, G.W.; Stirling, D.I. Immunomodulatory drugs inhibit expression of cyclooxygenase-2 from TNF-α, IL-1β, and LPS-stimulated human PBMC in a partially IL-10-dependent manner. Cell. Immunol. 2004, 230, 81–88. [Google Scholar] [CrossRef]
  207. Hideshima, T.; Chauhan, D.; Richardson, P.; Mitsiades, C.; Mitsiades, N.; Hayashi, T.; Munshi, N.; Dang, L.; Castro, A.; Palombella, V.; et al. NF-κB as a therapeutic target in multiple myeloma. J. Biol. Chem. 2002, 277, 16639–16647. [Google Scholar] [CrossRef] [PubMed]
  208. Keifer, J.A.; Guttridge, D.C.; Ashburner, B.P.; Baldwin, A.S. Inhibition of NF-κB Activity by Thalidomide through Suppression of IκB Kinase Activity. J. Biol. Chem. 2001, 276, 22382–22387. [Google Scholar] [CrossRef]
  209. Tachita, T.; Kinoshita, S.; Ri, M.; Aoki, S.; Asano, A.; Kanamori, T.; Yoshida, T.; Totani, H.; Ito, A.; Kusumoto, S.; et al. Expression, mutation, and methylation of cereblon-pathway genes at pre- and post-lenalidomide treatment in multiple myeloma. Cancer Sci. 2020, 111, 1333–1343. [Google Scholar] [CrossRef]
  210. Donovan, K.A.; An, J.; Nowak, R.P.; Yuan, J.C.; Fink, E.C.; Berry, B.C.; Ebert, B.L.; Fischer, E.S. Thalidomide promotes degradation of SALL4, a transcription factor implicated in Duane radial ray syndrome. Elife 2018, 7, e38430. [Google Scholar] [CrossRef]
  211. Aguiar, P.M.; de Mendonça Lima, T.; Colleoni, G.W.B.; Storpirtis, S. Efficacy and safety of bortezomib, thalidomide, and lenalidomide in multiple myeloma: An overview of systematic reviews with meta-analyses. Crit. Rev. Oncol. Hematol. 2017, 113, 195–212. [Google Scholar] [CrossRef]
  212. Nutt, S.L.; Hodgkin, P.D.; Tarlinton, D.M.; Corcoran, L.M. The generation of antibody-secreting plasma cells. Nat. Rev. Immunol. 2015, 15, 160–171. [Google Scholar] [CrossRef]
  213. Wang, H.; Morse, H.C.; Bolland, S. Transcriptional Control of Mature B Cell Fates. Trends Immunol. 2020, 41, 601–613. [Google Scholar] [CrossRef]
  214. Laidlaw, B.J.; Cyster, J.G. Transcriptional regulation of memory B cell differentiation. Nat. Rev. Immunol. 2021, 21, 209–220. [Google Scholar] [CrossRef]
  215. Ma, S.; Turetsky, A.; Trinh, L.; Lu, R. IFN Regulatory Factor 4 and 8 Promote Ig Light Chain κ Locus Activation in Pre-B Cell Development. J. Immunol. 2006, 177, 7898–7904. [Google Scholar] [CrossRef] [PubMed]
  216. Tellier, J.; Nutt, S.L. Plasma cells: The programming of an antibody-secreting machine. Eur. J. Immunol. 2019, 49, 30–37. [Google Scholar] [CrossRef] [PubMed]
  217. Shaffer, A.L.; Lin, K.I.; Kuo, T.C.; Yu, X.; Hurt, E.M.; Rosenwald, A.; Giltnane, J.M.; Yang, L.; Zhao, H.; Calame, K.; et al. Blimp-1 orchestrates plasma cell differentiation by extinguishing the mature B cell gene expression program. Immunity 2002, 17, 51–62. [Google Scholar] [CrossRef]
  218. Fricker, L.D. Proteasome Inhibitor Drugs. Annu. Rev. Pharmacol. Toxicol. 2020, 60, 457–476. [Google Scholar] [CrossRef]
  219. Ito, S. Proteasome inhibitors for the treatment of multiple myeloma. Cancers 2020, 12, 265. [Google Scholar] [CrossRef] [PubMed]
  220. Hershko, A. The ubiquitin system for protein degradation and some of its roles in the control of the cell division cycle. Cell Death Differ. 2005, 12, 1191–1197. [Google Scholar] [CrossRef]
  221. Adams, J.; Palombella, V.J.; Sausville, E.A.; Johnson, J.; Destree, A.; Lazarus, D.D.; Maas, J.; Pien, C.S.; Prakash, S.; Elliott, P.J. Proteasome inhibitors: A novel class of potent and effective antitumor agents. Cancer Res. 1999, 59, 2615–2622. [Google Scholar]
  222. Drexler, H.C.A. Activation of the cell death program by inhibition of proteasome function. Proc. Natl. Acad. Sci. USA 1997, 94, 855–860. [Google Scholar] [CrossRef] [PubMed]
  223. Gandolfi, S.; Laubach, J.P.; Hideshima, T.; Chauhan, D.; Anderson, K.C.; Richardson, P.G. The proteasome and proteasome inhibitors in multiple myeloma. Cancer Metastasis Rev. 2017, 36, 561–584. [Google Scholar] [CrossRef] [PubMed]
  224. Manasanch, E.E.; Orlowski, R.Z. Proteasome inhibitors in cancer therapy. Nat. Rev. Clin. Oncol. 2017, 14, 417–433. [Google Scholar] [CrossRef] [PubMed]
  225. Obeng, E.A.; Carlson, L.M.; Gutman, D.M.; Harrington, W.J.; Lee, K.P.; Boise, L.H. Proteasome inhibitors induce a terminal unfolded protein response in multiple myeloma cells. Blood 2006, 107, 4907–4916. [Google Scholar] [CrossRef]
  226. Meister, S.; Schubert, U.; Neubert, K.; Herrmann, K.; Burger, R.; Gramatzki, M.; Hahn, S.; Schreiber, S.; Wilhelm, S.; Herrmann, M.; et al. Extensive immunoglobulin production sensitizes myeloma cells for proteasome inhibition. Cancer Res. 2007, 67, 1783–1792. [Google Scholar] [CrossRef]
  227. Cavo, M.; Tacchetti, P.; Patriarca, F.; Petrucci, M.T.; Pantani, L.; Galli, M.; Di Raimondo, F.; Crippa, C.; Zamagni, E.; Palumbo, A.; et al. Bortezomib with thalidomide plus dexamethasone compared with thalidomide plus dexamethasone as induction therapy before, and consolidation therapy after, double autologous stem-cell transplantation in newly diagnosed multiple myeloma: A randomised phase 3. Lancet 2010, 376, 2075–2085. [Google Scholar] [CrossRef]
  228. Barlogie, B.; Anaissie, E.; Van Rhee, F.; Haessler, J.; Hollmig, K.; Pineda-Roman, M.; Cottler-Fox, M.; Mohiuddin, A.; Alsayed, Y.; Tricot, G.; et al. Incorporating bortezomib into upfront treatment for multiple myeloma: Early results of total therapy 3. Br. J. Haematol. 2007, 138, 176–185. [Google Scholar] [CrossRef]
  229. Richardson, P.G.; Xie, W.; Jagannath, S.; Jakubowiak, A.; Lonial, S.; Raje, N.S.; Alsina, M.; Ghobrial, I.M.; Schlossman, R.L.; Munshi, N.C.; et al. A phase 2 trial of lenalidomide, bortezomib, and dexamethasone in patients with relapsed and relapsed/refractory myeloma. Blood 2014, 123, 1461–1469. [Google Scholar] [CrossRef]
  230. Sonneveld, P.; Schmidt-Wolf, I.G.H.; Van Der Holt, B.; El Jarari, L.; Bertsch, U.; Salwender, H.; Zweegman, S.; Vellenga, E.; Broyl, A.; Blau, I.W.; et al. Bortezomib induction and maintenance treatment in patients with newly diagnosed multiple myeloma: Results of the randomized phase III HOVON-65/ GMMG-HD4 trial. J. Clin. Oncol. 2012, 30, 2946–2955. [Google Scholar] [CrossRef]
  231. Robak, P.; Robak, T. Bortezomib for the Treatment of Hematologic Malignancies: 15 Years Later. Drugs RD 2019, 19, 73–92. [Google Scholar] [CrossRef]
  232. Richardson, P.G.; Sonneveld, P.; Schuster, M.; Irwin, D.; Stadtmauer, E.; Facon, T.; Harousseau, J.L.; Ben-Yehuda, D.; Lonial, S.; Goldschmidt, H.; et al. Extended follow-up of a phase 3 trial in relapsed multiple myeloma: Final time-to-event results of the APEX trial. Blood 2007, 110, 3557–3560. [Google Scholar] [CrossRef] [PubMed]
  233. Richardson, P.G.; Barlogie, B.; Berenson, J.; Singhal, S.; Jagannath, S.; Irwin, D.; Rajkumar, S.V.; Srkalovic, G.; Alsina, M.; Alexanian, R.; et al. A Phase 2 Study of Bortezomib in Relapsed, Refractory Myeloma. N. Engl. J. Med. 2003, 348, 2609–2617. [Google Scholar] [CrossRef] [PubMed]
  234. Durie, B.G.M.; Hoering, A.; Abidi, M.H.; Rajkumar, S.V.; Epstein, J.; Kahanic, S.P.; Thakuri, M.; Reu, F.; Reynolds, C.M.; Sexton, R.; et al. Bortezomib with lenalidomide and dexamethasone versus lenalidomide and dexamethasone alone in patients with newly diagnosed myeloma without intent for immediate autologous stem-cell transplant (SWOG S0777): A randomised, open-label, phase 3 trial. Lancet 2017, 389, 519–527. [Google Scholar] [CrossRef]
  235. San Miguel, J.F.; Schlag, R.; Khuageva, N.K.; Dimopoulos, M.A.; Shpilberg, O.; Kropff, M.; Spicka, I.; Petrucci, M.T.; Palumbo, A.; Samoilova, O.S.; et al. Bortezomib plus Melphalan and Prednisone for Initial Treatment of Multiple Myeloma. N. Engl. J. Med. 2008, 359, 906–917. [Google Scholar] [CrossRef] [PubMed]
  236. Attal, M.; Lauwers-Cances, V.; Hulin, C.; Facon, T.; Caillot, D.; Escoffre, M.; Arnulf, B.; MACRO, M.; Belhadj, K.; Garderet, L.; et al. Autologous Transplantation for Multiple Myeloma in the Era of New Drugs: A Phase III Study of the Intergroupe Francophone Du Myelome (IFM/DFCI 2009 Trial). Blood 2015, 126, 391. [Google Scholar] [CrossRef]
  237. Goldschmidt, H.; Lokhorst, H.M.; Mai, E.K.; Van Der Holt, B.; Blau, I.W.; Zweegman, S.; Weisel, K.C.; Vellenga, E.; Pfreundschuh, M.; Kersten, M.J.; et al. Bortezomib before and after high-dose therapy in myeloma: Long-term results from the phase III HOVON-65/GMMG-HD4 trial. Leukemia 2018, 32, 383–390. [Google Scholar] [CrossRef]
  238. Richardson, P.G.; Sonneveld, P.; Schuster, M.W.; Irwin, D.; Stadtmauer, E.A.; Facon, T.; Harousseau, J.-L.; Ben-Yehuda, D.; Lonial, S.; Goldschmidt, H.; et al. Bortezomib or High-Dose Dexamethasone for Relapsed Multiple Myeloma. N. Engl. J. Med. 2005, 352, 2487–2498. [Google Scholar] [CrossRef]
  239. Stewart, A.K.; Rajkumar, S.V.; Dimopoulos, M.A.; Masszi, T.; Špička, I.; Oriol, A.; Hájek, R.; Rosiñol, L.; Siegel, D.S.; Mihaylov, G.G.; et al. Carfilzomib, Lenalidomide, and Dexamethasone for Relapsed Multiple Myeloma. N. Engl. J. Med. 2015, 372, 142–152. [Google Scholar] [CrossRef]
  240. Dimopoulos, M.A.; Moreau, P.; Palumbo, A.; Joshua, D.; Pour, L.; Hájek, R.; Facon, T.; Ludwig, H.; Oriol, A.; Goldschmidt, H.; et al. Carfilzomib and dexamethasone versus bortezomib and dexamethasone for patients with relapsed or refractory multiple myeloma (ENDEAVOR): And randomised, phase 3, open-label, multicentre study. Lancet Oncol. 2016, 17, 27–38. [Google Scholar] [CrossRef]
  241. Moreau, P.; Masszi, T.; Grzasko, N.; Bahlis, N.J.; Hansson, M.; Pour, L.; Sandhu, I.; Ganly, P.; Baker, B.W.; Jackson, S.R.; et al. Oral Ixazomib, Lenalidomide, and Dexamethasone for Multiple Myeloma. N. Engl. J. Med. 2016, 374, 1621–1634. [Google Scholar] [CrossRef]
  242. Dimopoulos, M.A.; Gay, F.; Schjesvold, F.; Beksac, M.; Hajek, R.; Weisel, K.C.; Goldschmidt, H.; Maisnar, V.; Moreau, P.; Min, C.K.; et al. Oral ixazomib maintenance following autologous stem cell transplantation (TOURMALINE-MM3): A double-blind, randomised, placebo-controlled phase 3 trial. Lancet 2019, 393, 253–264. [Google Scholar] [CrossRef]
  243. Hari, P.; Mateos, M.-V.; Abonour, R.; Knop, S.; Bensinger, W.; Ludwig, H.; Song, K.; Hajek, R.; Moreau, P.; Siegel, D.S.; et al. Efficacy and safety of carfilzomib regimens in multiple myeloma patients relapsing after autologous stem cell transplant: ASPIRE and ENDEAVOR outcomes. Leukemia 2017, 31, 2630–2641. [Google Scholar] [CrossRef] [PubMed]
  244. Dimopoulos, M.A.; Goldschmidt, H.; Niesvizky, R.; Joshua, D.; Chng, W.J.; Oriol, A.; Orlowski, R.Z.; Ludwig, H.; Facon, T.; Hajek, R.; et al. Carfilzomib or bortezomib in relapsed or refractory multiple myeloma (ENDEAVOR): An interim overall survival analysis of an open-label, randomised, phase 3 trial. Lancet Oncol. 2017, 18, 1327–1337. [Google Scholar] [CrossRef]
  245. Yarde, D.N.; Oliveira, V.; Mathews, L.; Wang, X.; Villagra, A.; Boulware, D.; Shain, K.H.; Hazlehurst, L.A.; Alsina, M.; Chen, D.T.; et al. Targeting the Fanconi anemia/BRCA pathway circumvents drug resistance in multiple myeloma. Cancer Res. 2009, 69, 9367–9375. [Google Scholar] [CrossRef] [PubMed]
  246. Neri, P.; Ren, L.; Gratton, K.; Stebner, E.; Johnson, J.; Klimowicz, A.; Duggan, P.; Tassone, P.; Mansoor, A.; Stewart, D.A.; et al. Bortezomib-induced “BRCAness” sensitizes multiple myeloma cells to PARP inhibitors. Blood 2011, 118, 6368–6379. [Google Scholar] [CrossRef]
  247. Jacquemont, C.; Taniguchi, T. Proteasome function is required for DNA damage response and fanconi anemia pathway activation. Cancer Res. 2007, 67, 7395–7405. [Google Scholar] [CrossRef]
  248. Murakawa, Y.; Sonoda, E.; Barber, L.J.; Zeng, W.; Yokomori, K.; Kimura, H.; Niimi, A.; Lehmann, A.; Guang, Y.Z.; Hochegger, H.; et al. Inhibitors of the proteasome suppress homologous DNA recombination in mammalian cells. Cancer Res. 2007, 67, 8536–8543. [Google Scholar] [CrossRef]
  249. Vrábel, D.; Pour, L.; Ševčíková, S. The impact of NF-κB signaling on pathogenesis and current treatment strategies in multiple myeloma. Blood Rev. 2019, 34, 56–66. [Google Scholar] [CrossRef]
  250. Hideshima, T.; Ikeda, H.; Chauhan, D.; Okawa, Y.; Raje, N.; Podar, K.; Mitsiades, C.; Munshi, N.C.; Richardson, P.G.; Carrasco, R.D.; et al. Bortezomib induces canonical nuclear factor-κB activation in multiple myeloma cells. Blood 2009, 114, 1046–1052. [Google Scholar] [CrossRef]
  251. Li, C.; Chen, S.; Yue, P.; Deng, X.; Lonial, S.; Khuri, F.R.; Sun, S.Y. Proteasome inhibitor PS-341 (Bortezomib) induces calpain-dependent IκBα degradation. J. Biol. Chem. 2010, 285, 16096–16104. [Google Scholar] [CrossRef]
  252. Xie, Z.; Bi, C.; Chooi, J.Y.; Chan, Z.L.; Mustafa, N.; Chng, W.J. MMSET regulates expression of IRF4 in t(4;14) myeloma and its silencing potentiates the effect of bortezomib. Leukemia 2015, 29, 2347–2354. [Google Scholar] [CrossRef]
  253. Marango, J.; Shimoyama, M.; Nishio, H.; Meyer, J.A.; Min, D.J.; Sirulnik, A.; Martinez-Martinez, Y.; Chesi, M.; Bergsagel, P.L.; Zhou, M.M.; et al. The MMSET protein is a histone methyltransferase with characteristics of a transcriptional corepressor. Blood 2008, 111, 3145–3154. [Google Scholar] [CrossRef] [PubMed]
  254. Xie, Z.; Chng, W.J. MMSET: Role and therapeutic opportunities in multiple myeloma. Biomed. Res. Int. 2014, 2014, 636514. [Google Scholar] [CrossRef]
  255. Barwick, B.G.; Gupta, V.A.; Vertino, P.M.; Boise, L.H. Cell of origin and genetic alterations in the pathogenesis of multiple myeloma. Front. Immunol. 2019, 10, 1121. [Google Scholar] [CrossRef] [PubMed]
  256. Imai, Y.; Hirano, M.; Kobayashi, M.; Futami, M.; Tojo, A. HDAC Inhibitors Exert Anti-Myeloma Effects through Multiple Modes of Action. Cancers 2019, 11, 475. [Google Scholar] [CrossRef] [PubMed]
  257. Hideshima, T.; Bradner, J.E.; Wong, J.; Chauhan, D.; Richardson, P.; Schreiber, S.L.; Anderson, K.C. Small-molecule inhibition of proteasome and aggresome function induces synergistic antitumor activity in multiple myeloma. Proc. Natl. Acad. Sci. USA 2005, 102, 8567–8572. [Google Scholar] [CrossRef]
  258. Pei, X.Y.; Dai, Y.; Grant, S. Synergistic induction of oxidative injury and apoptosis in human multiple myeloma cells by the proteasome inhibitor bortezomib and histone deacetylase inhibitors. Clin. Cancer Res. 2004, 10, 3839–3852. [Google Scholar] [CrossRef]
  259. Laubach, J.P.; Schjesvold, F.; Mariz, M.; Dimopoulos, M.A.; Lech-Maranda, E.; Spicka, I.; Hungria, V.T.M.; Shelekhova, T.; Abdo, A.; Jacobasch, L.; et al. Efficacy and safety of oral panobinostat plus subcutaneous bortezomib and oral dexamethasone in patients with relapsed or relapsed and refractory multiple myeloma (PANORAMA 3): An open-label, randomised, phase 2 study. Lancet Oncol. 2021, 22, 142–154. [Google Scholar] [CrossRef]
  260. San-Miguel, J.F.; Hungria, V.T.M.; Yoon, S.S.; Beksac, M.; Dimopoulos, M.A.; Elghandour, A.; Jedrzejczak, W.W.; Günther, A.; Nakorn, T.N.; Siritanaratkul, N.; et al. Overall survival of patients with relapsed multiple myeloma treated with panobinostat or placebo plus bortezomib and dexamethasone (the PANORAMA 1 trial): A randomised, placebo-controlled, phase 3 trial. Lancet Haematol. 2016, 3, e506–e515. [Google Scholar] [CrossRef]
  261. Wang, P.; Wang, Z.; Liu, J. Role of HDACs in normal and malignant hematopoiesis. Mol. Cancer 2020, 19, 5. [Google Scholar] [CrossRef]
  262. Boyault, C.; Zhang, Y.; Fritah, S.; Caron, C.; Gilquin, B.; So, H.K.; Garrido, C.; Yao, T.P.; Vourc’h, C.; Matthias, P.; et al. HDAC6 controls major cell response pathways to cytotoxic accumulation of protein aggregates. Genes Dev. 2007, 21, 2172–2181. [Google Scholar] [CrossRef] [PubMed]
  263. Tang, S.; Ma, D.; Cheng, B.; Fang, Q.; Kuang, X.; Yu, K.; Wang, W.; Hu, B.; Wang, J. Crucial role of HO-1/IRF4-dependent apoptosis induced by panobinostat and lenalidomide in multiple myeloma. Exp. Cell Res. 2018, 363, 196–207. [Google Scholar] [CrossRef]
  264. Bat-Erdene, A.; Miki, H.; Oda, A.; Nakamura, S.; Teramachi, J.; Amachi, R.; Tenshin, H.; Hiasa, M.; Iwasa, M.; Harada, T.; et al. Synergistic targeting of Sp1, a critical transcription factor for myeloma cell growth and survival, by panobinostat and proteasome inhibitors. Oncotarget 2016, 7, 79064–79075. [Google Scholar] [CrossRef] [PubMed]
  265. Ntanasis-Stathopoulos, I.; Gavriatopoulou, M.; Terpos, E. Antibody therapies for multiple myeloma. Expert Opin. Biol. Ther. 2020, 20, 295–303. [Google Scholar] [CrossRef]
  266. Dimopoulos, M.A.; San-Miguel, J.; Belch, A.; White, D.; Benboubker, L.; Cook, G.; Leiba, M.; Morton, J.; Joy Ho, P.; Kim, K.; et al. Daratumumab plus lenalidomide and dexamethasone versus lenalidomide and dexamethasone in relapsed or refractory multiple myeloma: Updated analysis of POLLUX. Haematologica 2018, 103, 2088–2096. [Google Scholar] [CrossRef] [PubMed]
  267. Spencer, A.; Lentzsch, S.; Weisel, K.; Avet-Loiseau, H.; Mark, T.M.; Spicka, I.; Masszi, T.; Lauri, B.; Levin, M.D.; Bosi, A.; et al. Daratumumab plus bortezomib and dexamethasone versus bortezomib and dexamethasone in relapsed or refractory multiple myeloma: Updated analysis of CASTOR. Haematologica 2018, 103, 2079–2087. [Google Scholar] [CrossRef] [PubMed]
  268. Attal, M.; Richardson, P.G.; Rajkumar, S.V.; San-Miguel, J.; Beksac, M.; Spicka, I.; Leleu, X.; Schjesvold, F.; Moreau, P.; Dimopoulos, M.A.; et al. Isatuximab plus pomalidomide and low-dose dexamethasone versus pomalidomide and low-dose dexamethasone in patients with relapsed and refractory multiple myeloma (ICARIA-MM): A randomised, multicentre, open-label, phase 3 study. Lancet 2019, 394, 2096–2107. [Google Scholar] [CrossRef]
  269. Van De Donk, N.W.C.J.; Richardson, P.G.; Malavasi, F. CD38 antibodies in multiple myeloma: Back to the future. Blood 2018, 131, 13–29. [Google Scholar] [CrossRef] [PubMed]
  270. Van De Donk, N.W.C.J.; Usmani, S.Z. CD38 antibodies in multiple myeloma: Mechanisms of action and modes of resistance. Front. Immunol. 2018, 9, 2134. [Google Scholar] [CrossRef] [PubMed]
  271. Covarrubias, A.J.; Perrone, R.; Grozio, A.; Verdin, E. NAD+ metabolism and its roles in cellular processes during ageing. Nat. Rev. Mol. Cell Biol. 2020, 22, 119–141. [Google Scholar] [CrossRef]
  272. D’agostino, M.; Innorcia, S.; Boccadoro, M.; Bringhen, S. Monoclonal antibodies to treat multiple myeloma: A dream come true. Int. J. Mol. Sci. 2020, 21, 8192. [Google Scholar] [CrossRef] [PubMed]
  273. Dima, D.; Dower, J.; Comenzo, R.L.; Varga, C. Evaluating daratumumab in the treatment of multiple myeloma: Safety, efficacy and place in therapy. Cancer Manag. Res. 2020, 12, 7891–7903. [Google Scholar] [CrossRef]
  274. Kennedy, B.E.; Sadek, M.; Elnenaei, M.O.; Reiman, A.; Gujar, S.A. Targeting NAD+ Synthesis to Potentiate CD38-Based Immunotherapy of Multiple Myeloma. Trends Cancer 2020, 6, 9–12. [Google Scholar] [CrossRef]
  275. Baum, N.; Fliegert, R.; Bauche, A.; Hambach, J.; Menzel, S.; Haag, F.; Bannas, P.; Koch-Nolte, F. Daratumumab and nanobody-based heavy chain antibodies inhibit the adpr cyclase but not the nad+ hydrolase activity of cd38-expressing multiple myeloma cells. Cancers 2021, 13, 76. [Google Scholar] [CrossRef] [PubMed]
  276. Katsyuba, E.; Romani, M.; Hofer, D.; Auwerx, J. NAD+ homeostasis in health and disease. Nat. Metab. 2020, 2, 9–31. [Google Scholar] [CrossRef] [PubMed]
  277. Saville, K.M.; Clark, J.; Wilk, A.; Rogers, G.D.; Andrews, J.F.; Koczor, C.A.; Sobol, R.W. NAD+-mediated regulation of mammalian base excision repair. DNA Repair (Amst) 2020, 93, 102930. [Google Scholar] [CrossRef] [PubMed]
  278. Campbell, K.S.; Cohen, A.D.; Pazina, T. Mechanisms of NK cell activation and clinical activity of the therapeutic SLAMF7 antibody, elotuzumab in multiple myeloma. Front. Immunol. 2018, 9, 2551. [Google Scholar] [CrossRef] [PubMed]
  279. Lonial, S.; Dimopoulos, M.; Palumbo, A.; White, D.; Grosicki, S.; Spicka, I.; Walter-Croneck, A.; Moreau, P.; Mateos, M.-V.; Magen, H.; et al. Elotuzumab Therapy for Relapsed or Refractory Multiple Myeloma. N. Engl. J. Med. 2015, 373, 621–631. [Google Scholar] [CrossRef]
  280. Dimopoulos, M.A.; Lonial, S.; White, D.; Moreau, P.; Weisel, K.; San-Miguel, J.; Shpilberg, O.; Grosicki, S.; Špička, I.; Walter-Croneck, A.; et al. Elotuzumab, lenalidomide, and dexamethasone in RRMM: Final overall survival results from the phase 3 randomized ELOQUENT-2 study. Blood Cancer J. 2020, 10, 91. [Google Scholar] [CrossRef]
  281. Dimopoulos, M.A.; Dytfeld, D.; Grosicki, S.; Moreau, P.; Takezako, N.; Hori, M.; Leleu, X.; LeBlanc, R.; Suzuki, K.; Raab, M.S.; et al. Elotuzumab plus Pomalidomide and Dexamethasone for Multiple Myeloma. N. Engl. J. Med. 2018, 379, 1811–1822. [Google Scholar] [CrossRef]
  282. Suzuki, A.; Kakugawa, S.; Miyoshi, M.; Hori, M.; Suzuki, K.; Furukawa, Y.; Ohta, K. Soluble SLAMF7 is a predictive biomarker for elotuzumab therapy. Leukemia 2020, 34, 3088–3090. [Google Scholar] [CrossRef] [PubMed]
  283. Trudel, S.; Moreau, P.; Touzeau, C. Update on elotuzumab for the treatment of relapsed/refractory multiple myeloma: Patients’ selection and perspective. Onco. Targets. Ther. 2019, 12, 5813–5822. [Google Scholar] [CrossRef]
  284. Guo, H.; Cruz-Munoz, M.-E.; Wu, N.; Robbins, M.; Veillette, A. Immune Cell Inhibition by SLAMF7 Is Mediated by a Mechanism Requiring Src Kinases, CD45, and SHIP-1 That Is Defective in Multiple Myeloma Cells. Mol. Cell. Biol. 2015, 35, 41–51. [Google Scholar] [CrossRef]
  285. Balasa, B.; Yun, R.; Belmar, N.A.; Fox, M.; Chao, D.T.; Robbins, M.D.; Starling, G.C.; Rice, A.G. Elotuzumab enhances natural killer cell activation and myeloma cell killing through interleukin-2 and TNF-α pathways. Cancer Immunol. Immunother. 2015, 64, 61–73. [Google Scholar] [CrossRef] [PubMed]
  286. Xie, Z.; Gunaratne, J.; Cheong, L.L.; Liu, S.C.; Koh, T.L.; Huang, G.; Blackstock, W.P.; Chng, W.J. Plasma membrane proteomics identifies biomarkers associated with MMSET overexpression in t(4;14) multiple myeloma. Oncotarget 2013, 4, 1008–1018. [Google Scholar] [CrossRef]
  287. Gabellini, C.; Trisciuoglio, D.; Bufalo, D. Del Non-canonical roles of Bcl-2 and Bcl-xL proteins: Relevance of BH4 domain. Carcinogenesis 2017, 38, 579–587. [Google Scholar] [CrossRef] [PubMed]
  288. Jamil, S.; Stoica, C.; Hackett, T.L.; Duronio, V. MCL-1 localizes to sites of DNA damage and regulates DNA damage response. Cell Cycle 2010, 9, 2915–2927. [Google Scholar] [CrossRef]
  289. Mattoo, A.R.; Pandita, R.K.; Chakraborty, S.; Charaka, V.; Mujoo, K.; Hunt, C.R.; Pandita, T.K. MCL-1 Depletion Impairs DNA Double-Strand Break Repair and Reinitiation of Stalled DNA Replication Forks. Mol. Cell. Biol. 2017, 37, e00535-16. [Google Scholar] [CrossRef]
  290. Gong, J.N.; Khong, T.; Segal, D.; Yao, Y.; Riffkin, C.D.; Garnier, J.M.; Lin Khaw, S.; Lessene, G.; Spencer, A.; Herold, M.J.; et al. Hierarchy for targeting prosurvival BCL2 family proteins in multiple myeloma: Pivotal role of MCL1. Blood 2016, 128, 1834–1844. [Google Scholar] [CrossRef]
  291. Tron, A.E.; Belmonte, M.A.; Adam, A.; Aquila, B.M.; Boise, L.H.; Chiarparin, E.; Cidado, J.; Embrey, K.J.; Gangl, E.; Gibbons, F.D.; et al. Discovery of Mcl-1-specific inhibitor AZD5991 and preclinical activity in multiple myeloma and acute myeloid leukemia. Nat. Commun. 2018, 9, 5341. [Google Scholar] [CrossRef]
  292. Kotschy, A.; Szlavik, Z.; Murray, J.; Davidson, J.; Maragno, A.L.; Le Toumelin-Braizat, G.; Chanrion, M.; Kelly, G.L.; Gong, J.N.; Moujalled, D.M.; et al. The MCL1 inhibitor S63845 is tolerable and effective in diverse cancer models. Nature 2016, 538, 477–482. [Google Scholar] [CrossRef]
  293. Touzeau, C.; Maciag, P.; Amiot, M.; Moreau, P. Targeting Bcl-2 for the treatment of multiple myeloma. Leukemia 2018, 32, 1899–1907. [Google Scholar] [CrossRef] [PubMed]
  294. Paner, A.; Patel, P.; Dhakal, B. The evolving role of translocation t(11;14) in the biology, prognosis, and management of multiple myeloma. Blood Rev. 2020, 41, 100643. [Google Scholar] [CrossRef] [PubMed]
  295. Punnoose, E.A.; Leverson, J.D.; Peale, F.; Boghaert, E.R.; Belmont, L.D.; Tan, N.; Young, A.; Mitten, M.; Ingalla, E.; Darbonne, W.C.; et al. Expression profile of BCL-2, BCL-XL, and MCL-1 predicts pharmacological response to the BCL-2 selective antagonist venetoclax in multiple myeloma models. Mol. Cancer Ther. 2016, 15, 1132–1144. [Google Scholar] [CrossRef]
  296. Chauhan, D.; Auclair, D.; Robinson, E.K.; Hideshima, T.; Li, G.; Podar, K.; Gupta, D.; Richardson, P.; Schlossma, R.L.; Krett, N.; et al. Identification of genes regulated by Dexamethasone in multiple myeloma cells using oligonucleotide arrays. Oncogene 2002, 21, 1346–1358. [Google Scholar] [CrossRef]
  297. Burington, B.; Barlogie, B.; Zhan, F.; Crowley, J.; Shaughnessy, J.D. Tumor cell gene expression changes following short-term in vivo exposure to single agent chemotherapeutics are related to survival in multiple myeloma. Clin. Cancer Res. 2008, 14, 4821. [Google Scholar] [CrossRef]
  298. Cannon-Albright, L.A.; Thomas, A.; Goldgar, D.E.; Gholami, K.; Rowe, K.; Jacobsen, M.; McWhorter, W.P.; Skolnick, M.H. Familiality of Cancer in Utah. Cancer Res. 1994, 54, 2378–2385. [Google Scholar] [PubMed]
  299. Landgren, O.; Linet, M.S.; McMaster, M.L.; Gridley, G.; Hemminki, K.; Goldin, L.R. Familial characteristics of autoimmune and hematologic disorders in 8,406 multiple myeloma patients: A population-based case-control study. Int. J. Cancer 2006, 118, 3095–3098. [Google Scholar] [CrossRef]
  300. Albright, F.; Teerlink, C.; Werner, T.L.; Cannon-Albright, L.A. Significant evidence for a heritable contribution to cancer predisposition: A review of cancer familiality by site. BMC Cancer 2012, 12, 138. [Google Scholar] [CrossRef]
  301. Schinasi, L.H.; Brown, E.E.; Camp, N.J.; Wang, S.S.; Hofmann, J.N.; Chiu, B.C.; Miligi, L.; Beane Freeman, L.E.; de Sanjose, S.; Bernstein, L.; et al. Multiple myeloma and family history of lymphohaematopoietic cancers: Results from the International Multiple Myeloma Consortium. Br. J. Haematol. 2016, 175, 87–101. [Google Scholar] [CrossRef]
  302. Landgren, O.; Kristinsson, S.Y.; Goldin, L.R.; Caporaso, N.E.; Blimark, C.; Mellqvist, U.H.; Wahlin, A.; Bjorkholm, M.; Turesson, I. Risk of plasma cell and lymphoproliferative disorders among 14 621 first-degree relatives of 4458 patients with monoclonal gammopathy of undetermined significance in Sweden. Blood 2009, 114, 791–795. [Google Scholar] [CrossRef]
  303. Greenberg, A.J.; Vincent Rajkumar, S.; Larson, D.R.; Dispenzieri, A.; Therneau, T.M.; Colby, C.L.; Phelps, T.K.; Kumar, S.K.; Katzmann, J.A.; Kyle, R.A.; et al. Increased prevalence of light chain monoclonal gammopathy of undetermined significance (LC-MGUS) in first-degree relatives of individuals with multiple myeloma. Br. J. Haematol. 2012, 157, 472–475. [Google Scholar] [CrossRef] [PubMed]
  304. Vachon, C.M.; Kyle, R.A.; Therneau, T.M.; Foreman, B.J.; Larson, D.R.; Colby, C.L.; Phelps, T.K.; Dispenzieri, A.; Kumar, S.K.; Katzmann, J.A.; et al. Increased risk of monoclonal gammopathy in first-degree relatives of patients with multiple myeloma or monoclonal gammopathy of undetermined significance. Blood 2009, 114, 785–790. [Google Scholar] [CrossRef]
  305. Wei, X.; Calvo-Vidal, M.N.; Chen, S.; Wu, G.; Revuelta, M.V.; Sun, J.; Zhang, J.; Walsh, M.F.; Nichols, K.E.; Joseph, V.; et al. Germline lysine-specific demethylase 1 (lsd1/kdm1a) mutations confer susceptibility to multiple myeloma. Cancer Res. 2018, 78, 2747–2759. [Google Scholar] [CrossRef] [PubMed]
  306. Saleque, S.; Kim, J.; Rooke, H.M.; Orkin, S.H. Epigenetic Regulation of Hematopoietic Differentiation by Gfi-1 and Gfi-1b Is Mediated by the Cofactors CoREST and LSD1. Mol. Cell 2007, 27, 562–572. [Google Scholar] [CrossRef]
  307. De Smedt, E.; Lui, H.; Maes, K.; De Veirman, K.; Menu, E.; Vanderkerken, K.; De Bruyne, E. The epigenome in multiple myeloma: Impact on tumor cell plasticity and drug response. Front. Oncol. 2018, 8, 566. [Google Scholar] [CrossRef]
  308. Shi, Y.J.; Matson, C.; Lan, F.; Iwase, S.; Baba, T.; Shi, Y. Regulation of LSD1 histone demethylase activity by its associated factors. Mol. Cell 2005, 19, 857–864. [Google Scholar] [CrossRef] [PubMed]
  309. Yang, Y.; Huang, W.; Qiu, R.; Liu, R.; Zeng, Y.; Gao, J.; Zheng, Y.; Hou, Y.; Wang, S.; Yu, W.; et al. LSD1 coordinates with the SIN3A/HDAC complex and maintains sensitivity to chemotherapy in breast cancer. J. Mol. Cell Biol. 2018, 10, 285–301. [Google Scholar] [CrossRef]
  310. Gu, F.; Lin, Y.; Wang, Z.; Wu, X.; Ye, Z.; Wang, Y.; Lan, H. Biological roles of LSD1 beyond its demethylase activity. Cell. Mol. Life Sci. 2020, 77, 3341–3350. [Google Scholar] [CrossRef]
  311. Hyun, K.; Jeon, J.; Park, K.; Kim, J. Writing, erasing and reading histone lysine methylations. Exp. Mol. Med. 2017, 49, 324. [Google Scholar] [CrossRef]
  312. Nanou, A.; Toumpeki, C.; Lavigne, M.D.; Lazou, V.; Demmers, J.; Paparountas, T.; Thanos, D.; Katsantoni, E. The dual role of LSD1 and HDAC3 in STAT5-dependent transcription is determined by protein interactions, binding affinities, motifs and genomic positions. Nucleic Acids Res. 2017, 45, 142–154. [Google Scholar] [CrossRef]
  313. Karakaidos, P.; Verigos, J.; Magklara, A. Lsd1/kdm1a, a gate-keeper of cancer stemness and a promising therapeutic target. Cancers 2019, 11, 1821. [Google Scholar] [CrossRef] [PubMed]
  314. Majello, B.; Gorini, F.; Saccà, C.D.; Amente, S. Expanding the role of the histone lysine-specific demethylase lsd1 in cancer. Cancers 2019, 11, 324. [Google Scholar] [CrossRef]
  315. Hatzi, K.; Geng, H.; Doane, A.S.; Meydan, C.; LaRiviere, R.; Cardenas, M.; Duy, C.; Shen, H.; Vidal, M.N.C.; Baslan, T.; et al. Histone demethylase LSD1 is required for germinal center formation and BCL6-driven lymphomagenesis. Nat. Immunol. 2019, 20, 86–96. [Google Scholar] [CrossRef]
  316. Roco, J.A.; Mesin, L.; Binder, S.C.; Nefzger, C.; Gonzalez-Figueroa, P.; Canete, P.F.; Ellyard, J.; Shen, Q.; Robert, P.A.; Cappello, J.; et al. Class-Switch Recombination Occurs Infrequently in Germinal Centers. Immunity 2019, 51, 337–350.e7. [Google Scholar] [CrossRef] [PubMed]
  317. Huang, J.; Sengupta, R.; Espejo, A.B.; Lee, M.G.; Dorsey, J.A.; Richter, M.; Opravil, S.; Shiekhattar, R.; Bedford, M.T.; Jenuwein, T.; et al. p53 is regulated by the lysine demethylase LSD1. Nature 2007, 449, 105–108. [Google Scholar] [CrossRef]
  318. Escoubet-Lozach, L.; Lin, I.L.; Jensen-Pergakes, K.; Brady, H.A.; Gandhi, A.K.; Schafer, P.H.; Muller, G.W.; Worland, P.J.; Chan, K.W.H.; Verhelle, D. Pomalidomide and lenalidomide induce p21WAF-1 expression in both lymphoma and multiple myeloma through a LSD1-mediated epigenetic mechanism. Cancer Res. 2009, 69, 7347–7356. [Google Scholar] [CrossRef]
  319. Waller, R.G.; Darlington, T.M.; Wei, X.; Madsen, M.J.; Thomas, A.; Curtin, K.; Coon, H.; Rajamanickam, V.; Musinsky, J.; Jayabalan, D.; et al. Novel pedigree analysis implicates DNA repair and chromatin remodeling in multiple myeloma risk. PLOS Genet. 2018, 14, e1007111. [Google Scholar] [CrossRef]
  320. Masliah-Planchon, J.; Bièche, I.; Guinebretière, J.-M.; Bourdeaut, F.; Delattre, O. SWI/SNF Chromatin Remodeling and Human Malignancies. Annu. Rev. Pathol. Mech. Dis. 2015, 10, 145–171. [Google Scholar] [CrossRef]
  321. Mathur, R. ARID1A loss in cancer: Towards a mechanistic understanding. Pharmacol. Ther. 2018, 190, 15–23. [Google Scholar] [CrossRef] [PubMed]
  322. Yamamoto, H.; Imai, K. Microsatellite instability: An update. Arch. Toxicol. 2015, 89, 899–921. [Google Scholar] [CrossRef]
  323. Dykhuizen, E.C.; Hargreaves, D.C.; Miller, E.L.; Cui, K.; Korshunov, A.; Kool, M.; Pfister, S.; Cho, Y.J.; Zhao, K.; Crabtree, G.R. BAF complexes facilitate decatenation of DNA by topoisomerase IIα. Nature 2013, 497, 624–627. [Google Scholar] [CrossRef]
  324. Trizzino, M.; Barbieri, E.; Petracovici, A.; Wu, S.; Welsh, S.A.; Owens, T.A.; Licciulli, S.; Zhang, R.; Gardini, A. The Tumor Suppressor ARID1A Controls Global Transcription via Pausing of RNA Polymerase II. Cell Rep. 2018, 23, 3933–3945. [Google Scholar] [CrossRef] [PubMed]
  325. Perez-Oliva, A.B.; Lachaud, C.; Szyniarowski, P.; Muñoz, I.; Macartney, T.; Hickson, I.; Rouse, J.; Alessi, D.R. USP 45 deubiquitylase controls ERCC 1– XPF endonuclease-mediated DNA damage responses. EMBO J. 2015, 34, 326–343. [Google Scholar] [CrossRef]
  326. Schrader, C.E.; Vardo, J.; Linehan, E.; Twarog, M.Z.; Niedernhofer, L.J.; Hoeijmakers, J.H.J.; Stavnezer, J. Deletion of the nucleotide excision repair gene Ercc1 reduces immunoglobulin class switching and alters mutations near switch recombination junctions. J. Exp. Med. 2004, 200, 321–330. [Google Scholar] [CrossRef] [PubMed]
  327. Weißbach, S.; Langer, C.; Puppe, B.; Nedeva, T.; Bach, E.; Kull, M.; Bargou, R.; Einsele, H.; Rosenwald, A.; Knop, S.; et al. The molecular spectrum and clinical impact of DIS3 mutations in multiple myeloma. Br. J. Haematol. 2015, 169, 57–70. [Google Scholar] [CrossRef]
  328. Bolli, N.; Biancon, G.; Moarii, M.; Gimondi, S.; Li, Y.; de Philippis, C.; Maura, F.; Sathiaseelan, V.; Tai, Y.T.Y.T.; Mudie, L.; et al. Analysis of the genomic landscape of multiple myeloma highlights novel prognostic markers and disease subgroups. Leukemia 2018, 32, 2604–2616. [Google Scholar] [CrossRef] [PubMed]
  329. Ruiz-Heredia, Y.; Sánchez-Vega, B.; Onecha, E.; Barrio, S.; Alonso, R.; Martínez-ávila, J.C.; Cuenca, I.; Agirre, X.; Braggio, E.; Hernández, M.T.; et al. Mutational screening of newly diagnosed multiple myeloma patients by deep targeted sequencing. Haematologica 2018, 103, e544–e548. [Google Scholar] [CrossRef]
  330. Pertesi, M.; Vallée, M.; Wei, X.; Revuelta, M.V.; Galia, P.; Demangel, D.; Oliver, J.; Foll, M.; Chen, S.; Perrial, E.; et al. Exome sequencing identifies germline variants in DIS3 in familial multiple myeloma. Leukemia 2019, 33, 2324–2330. [Google Scholar] [CrossRef]
  331. Laffleur, B.; Basu, U.; Lim, J. RNA Exosome and Non-coding RNA-Coupled Mechanisms in AID-Mediated Genomic Alterations. J. Mol. Biol. 2017, 429, 3230–3241. [Google Scholar] [CrossRef]
  332. Basu, U.; Meng, F.L.; Keim, C.; Grinstein, V.; Pefanis, E.; Eccleston, J.; Zhang, T.; Myers, D.; Wasserman, C.R.; Wesemann, D.R.; et al. The RNA exosome targets the AID cytidine deaminase to both strands of transcribed duplex DNA substrates. Cell 2011, 144, 353–363. [Google Scholar] [CrossRef]
  333. Bolli, N.; Barcella, M.; Salvi, E.; D’Avila, F.; Vendramin, A.; De Philippis, C.; Munshi, N.C.; Avet-Loiseau, H.; Campbell, P.J.; Mussetti, A.; et al. Next-generation sequencing of a family with a high penetrance of monoclonal gammopathies for the identification of candidate risk alleles. Cancer 2017, 123, 3701–3708. [Google Scholar] [CrossRef]
  334. Dilworth, D.; Liu, L.; Stewart, A.K.; Berenson, J.R.; Lassam, N.; Hogg, D. Germline CDKN2A mutation implicated in predisposition to multiple myeloma. Blood 2000, 95, 1869–1871. [Google Scholar] [CrossRef]
  335. Shah, V.; Boyd, K.D.; Houlston, R.S.; Kaiser, M.F. Constitutional mutation in CDKN2A is associated with long term survivorship in multiple myeloma: A case report. BMC Cancer 2017, 17, 718. [Google Scholar] [CrossRef]
  336. Rayess, H.; Wang, M.B.; Srivatsan, E.S. Cellular senescence and tumor suppressor gene p16. Int. J. Cancer 2012, 130, 1715–1725. [Google Scholar] [CrossRef] [PubMed]
  337. He, S.; Sharpless, N.E. Senescence in Health and Disease. Cell 2017, 169, 1000–1011. [Google Scholar] [CrossRef] [PubMed]
  338. Sherr, C.J. The INK4a/ARF network in tumour suppression. Nat. Rev. Mol. Cell Biol. 2001, 2, 731–737. [Google Scholar] [CrossRef] [PubMed]
  339. Tessoulin, B.; Eveillard, M.; Lok, A.; Chiron, D.; Moreau, P.; Amiot, M.; Moreau-Aubry, A.; Le Gouill, S.; Pellat-Deceunynck, C. p53 dysregulation in B-cell malignancies: More than a single gene in the pathway to hell. Blood Rev. 2017, 31, 251–259. [Google Scholar] [CrossRef]
  340. Stanganelli, C.; Arbelbide, J.; Fantl, D.B.; Corrado, C.; Slavutsky, I. DNA methylation analysis of tumor suppressor genes in monoclonal gammopathy of undetermined significance. Ann. Hematol. 2010, 89, 191–199. [Google Scholar] [CrossRef] [PubMed]
  341. Ng, M.H.L.; Chung, Y.F.; Lo, K.W.; Wickham, N.W.R.; Lee, J.C.K.; Huang, D.P. Frequent hypermethylation of p16 and p15 genes in multiple myeloma. Blood 1997, 89, 2500–2506. [Google Scholar] [CrossRef] [PubMed]
  342. Taniguchi, T.; Chikatsu, N.; Takahashi, S.; Fujita, A.; Uchimaru, K.; Asano, S.; Fujita, T.; Motokura, T. Expression of p16(INK4A) and p14(ARF) in hematological malignancies. Leukemia 1999, 13, 1760–1769. [Google Scholar] [CrossRef]
  343. Chubb, D.; Weinhold, N.; Broderick, P.; Chen, B.; Johnson, D.C.; Försti, A.; Vijayakrishnan, J.; Migliorini, G.; Dobbins, S.E.; Holroyd, A.; et al. Common variation at 3q26.2, 6p21.33, 17p11.2 and 22q13.1 influences multiple myeloma risk. Nat. Genet. 2013, 45, 1221–1225. [Google Scholar] [CrossRef]
  344. Mitchell, J.S.; Li, N.; Weinhold, N.; Försti, A.; Ali, M.; Van Duin, M.; Thorleifsson, G.; Johnson, D.C.; Chen, B.; Halvarsson, B.M.; et al. Genome-wide association study identifies multiple susceptibility loci for multiple myeloma. Nat. Commun. 2016, 7, 12050. [Google Scholar] [CrossRef]
  345. Went, M.; Sud, A.; Försti, A.; Halvarsson, B.M.; Weinhold, N.; Kimber, S.; van Duin, M.; Thorleifsson, G.; Holroyd, A.; Johnson, D.C.; et al. Identification of multiple risk loci and regulatory mechanisms influencing susceptibility to multiple myeloma. Nat. Commun. 2018, 9, 3707. [Google Scholar] [CrossRef] [PubMed]
  346. Pertesi, M.; Went, M.; Hansson, M.; Hemminki, K.; Houlston, R.S.; Nilsson, B. Genetic predisposition for multiple myeloma. Leukemia 2020, 34, 697–708. [Google Scholar] [CrossRef]
  347. Janz, S.; Zhan, F.; Sun, F.; Cheng, Y.; Pisano, M.; Yang, Y.; Goldschmidt, H.; Hari, P. Germline risk contribution to genomic instability in multiple myeloma. Front. Genet. 2019, 10, 424. [Google Scholar] [CrossRef]
  348. Broderick, P.; Chubb, D.; Johnson, D.C.; Weinhold, N.; Försti, A.; Lloyd, A.; Olver, B.; Ma, Y.P.; Dobbins, S.E.; Walker, B.A.; et al. Common variation at 3p22.1 and 7p15.3 influences multiple myeloma risk. Nat. Genet. 2012, 44, 58–61. [Google Scholar] [CrossRef] [PubMed]
  349. Weinhold, N.; Johnson, D.C.; Chubb, D.; Chen, B.; Försti, A.; Hosking, F.J.; Broderick, P.; Ma, Y.P.; Dobbins, S.E.; Hose, D.; et al. The CCND1 c.870G>A polymorphism is a risk factor for t(11;14)(q13;q32) multiple myeloma. Nat. Genet. 2013, 45, 522–525. [Google Scholar] [CrossRef]
  350. Martino, A.; Campa, D.; Jamroziak, K.; Reis, R.M.; Sainz, J.; Buda, G.; García-Sanz, R.; Lesueur, F.; Marques, H.; Moreno, V.; et al. Impact of polymorphic variation at 7p15.3, 3p22.1 and 2p23.3 loci on risk of multiple myeloma. Br. J. Haematol. 2012, 158, 805–809. [Google Scholar] [CrossRef]
  351. Erickson, S.W.; Raj, V.R.; Stephens, O.W.; Dhakal, I.; Chavan, S.S.; Sanathkumar, N.; Coleman, E.A.; Lee, J.Y.; Goodwin, J.A.; Apewokin, S.; et al. Genome-wide scan identifies variant in 2q12.3 associated with risk for multiple myeloma. Blood 2014, 124, 2001–2003. [Google Scholar] [CrossRef] [PubMed]
  352. Weinhold, N.; Johnson, D.C.; Rawstron, A.C.; Försti, A.; Doughty, C.; Vijayakrishnan, J.; Broderick, P.; Dahir, N.B.; Begum, D.B.; Hosking, F.J.; et al. Inherited genetic susceptibility to monoclonal gammopathy of unknown significance. Blood 2014, 123, 2513–2517. [Google Scholar] [CrossRef] [PubMed]
  353. Thomsen, H.; Campo, C.; Weinhold, N.; da Silva Filho, M.I.; Pour, L.; Gregora, E.; Vodicka, P.; Vodickova, L.; Hoffmann, P.; Nöthen, M.M.; et al. Genomewide association study on monoclonal gammopathy of unknown significance (MGUS). Eur. J. Haematol. 2017, 99, 70–79. [Google Scholar] [CrossRef] [PubMed]
  354. Went, M.; Kinnersley, B.; Sud, A.; Johnson, D.C.; Weinhold, N.; Försti, A.; van Duin, M.; Orlando, G.; Mitchell, J.S.; Kuiper, R.; et al. Transcriptome-wide association study of multiple myeloma identifies candidate susceptibility genes. Hum. Genomics 2019, 13, 37. [Google Scholar] [CrossRef] [PubMed]
  355. Clay-Gilmour, A.I.; Hildebrandt, M.A.T.; Brown, E.E.; Hofmann, J.N.; Spinelli, J.J.; Giles, G.G.; Cozen, W.; Bhatti, P.; Wu, X.; Waller, R.G.; et al. Coinherited genetics of multiple myeloma and its precursor, monoclonal gammopathy of undetermined significance. Blood Adv. 2020, 4, 2789–2797. [Google Scholar] [CrossRef] [PubMed]
  356. Piunti, A.; Shilatifard, A. Epigenetic balance of gene expression by polycomb and compass families. Science 2016, 352. [Google Scholar] [CrossRef]
  357. Barbour, H.; Daou, S.; Hendzel, M.; Affar, E.B. Polycomb group-mediated histone H2A monoubiquitination in epigenome regulation and nuclear processes. Nat. Commun. 2020, 11, 5947. [Google Scholar] [CrossRef] [PubMed]
  358. Scott, C.L.; Gil, J.; Hernando, E.; Teruya-Feldstein, J.; Narita, M.; Martínez, D.; Visakorpi, T.; Mu, D.; Cordon-Cardo, C.; Peters, G.; et al. Role of the chromobox protein CBX7 in lymphomagenesis. Proc. Natl. Acad. Sci. USA 2007, 104, 5389–5394. [Google Scholar] [CrossRef]
  359. Hu, D.; Shilatifard, A. Epigenetics of hematopoiesis and hematological malignancies. Genes Dev. 2016, 30, 2021–2041. [Google Scholar] [CrossRef]
  360. Di Carlo, V.; Mocavini, I.; Di Croce, L. Polycomb complexes in normal and malignant hematopoiesis. J. Cell Biol. 2019, 218, 55–69. [Google Scholar] [CrossRef]
  361. Ginjala, V.; Nacerddine, K.; Kulkarni, A.; Oza, J.; Hill, S.J.; Yao, M.; Citterio, E.; van Lohuizen, M.; Ganesan, S. BMI1 Is Recruited to DNA Breaks and Contributes to DNA Damage-Induced H2A Ubiquitination and Repair. Mol. Cell. Biol. 2011, 31, 1972–1982. [Google Scholar] [CrossRef]
  362. Ismail, H.; Andrin, C.; McDonald, D.; Hendzel, M.J. BMI1-mediated histone ubiquitylation promotes DNA double-strand break repair. J. Cell Biol. 2010, 191, 45–60. [Google Scholar] [CrossRef] [PubMed]
  363. Jagani, Z.; Wiederschain, D.; Loo, A.; He, D.; Mosher, R.; Fordjour, P.; Monahan, J.; Morrissey, M.; Yao, Y.M.; Lengauer, C.; et al. The polycomb group protein Bmi-1 is essential for the growth of multiple myeloma cells. Cancer Res. 2010, 70, 5528–5538. [Google Scholar] [CrossRef]
  364. Wu, S.Q.; Xu, Z.Z.; Niu, W.Y.; Huang, H.B.; Zhan, R. shRNA-mediated Bmi-1 silencing sensitizes multiple myeloma cells to bortezomib. Int. J. Mol. Med. 2014, 34, 616–623. [Google Scholar] [CrossRef] [PubMed]
  365. Laugesen, A.; Højfeldt, J.W.; Helin, K. Role of the polycomb repressive complex 2 (PRC2) in transcriptional regulation and cancer. Cold Spring Harb. Perspect. Med. 2016, 6, a026575. [Google Scholar] [CrossRef] [PubMed]
  366. Bracken, A.P.; Kleine-Kohlbrecher, D.; Dietrich, N.; Pasini, D.; Gargiulo, G.; Beekman, C.; Theilgaard-Mönch, K.; Minucci, S.; Porse, B.T.; Marine, J.C.; et al. The Polycomb group proteins bind throughout the INK4A-ARF locus and are disassociated in senescent cells. Genes Dev. 2007, 21, 525–530. [Google Scholar] [CrossRef] [PubMed]
  367. Agherbi, H.; Gaussmann-Wenger, A.; Verthuy, C.; Chasson, L.; Serrano, M.; Djabali, M. Polycomb mediated epigenetic silencing and replication timing at the INK4a/ARF locus during senescence. PLoS ONE 2009, 4, e5622. [Google Scholar] [CrossRef]
  368. Rand, K.A.; Song, C.; Dean, E.; Serie, D.J.; Curtin, K.; Sheng, X.; Hu, D.; Huff, C.A.; Bernal-Mizrachi, L.; Tomasson, H.M.; et al. A meta-analysis of multiple myeloma risk regions in African and European ancestry populations identifies putatively functional loci. Cancer Epidemiol. Biomarkers Prev. 2016, 25, 1609–1618. [Google Scholar] [CrossRef]
  369. Weinhold, N.; Meissner, T.; Johnson, D.C.; Seckinger, A.; Moreaux, J.; Försti, A.; Chen, B.; Nickel, J.; Chubb, D.; Rawstron, A.C.; et al. The 7p15.3 (rs4487645) association for multiple myeloma shows strong allele-specific regulation of the myc-interacting gene cdca7l in malignant plasma cells. Haematologica 2015, 100, e110–e113. [Google Scholar] [CrossRef] [PubMed]
  370. Li, N.; Johnson, D.C.; Weinhold, N.; Studd, J.B.; Orlando, G.; Mirabella, F.; Mitchell, J.S.; Meissner, T.; Kaiser, M.; Goldschmidt, H.; et al. Multiple myeloma risk variant at 7p15.3 creates an IRF4-binding site and interferes with CDCA7L expression. Nat. Commun. 2016, 7, 13656. [Google Scholar] [CrossRef]
  371. Ji, Q.K.; Ma, J.W.; Liu, R.H.; Li, X.S.; Shen, F.Z.; Huang, L.Y.; Hui, L.; Ma, Y.J.; Jin, B.Z. CDCA7L promotes glioma proliferation by targeting CCND1 and predicts an unfavorable prognosis. Mol. Med. Rep. 2019, 20, 1149–1156. [Google Scholar] [CrossRef]
  372. Hop, P.J.; Luijk, R.; Daxinger, L.; Iterson, M.V.; Dekkers, K.F.; Jansen, R.; Meurs, J.B.J.V.; Hoen, P.A.C.T.; Ikram, M.A.; Greevenbroek, M.M.J.V.; et al. Genome-wide identification of genes regulating DNA methylation using genetic anchors for causal inference. Genome Biol. 2020, 21, 220. [Google Scholar] [CrossRef]
  373. Karki, K.; Harishchandra, S.; Safe, S. Bortezomib targets SP transcription factors in cancer cells s. Mol. Pharmacol. 2018, 94, 1187–1196. [Google Scholar] [CrossRef]
  374. Winkelmann, R.; Sandrock, L.; Kirberg, J.; Jäck, H.M.; Schuh, W. KLF2- a negative regulator of pre-B cell clonal expansion and B cell activation. PLoS ONE 2014, 9, e97953. [Google Scholar]
  375. Salama, Y.; Heida, A.H.; Yokoyama, K.; Takahashi, S.; Hattori, K.; Heissig, B. The EGFL7-ITGB3-KLF2 axis enhances survival of multiple myeloma in preclinical models. Blood Adv. 2020, 4, 1021–1037. [Google Scholar] [CrossRef] [PubMed]
  376. Ali, M.; Ajore, R.; Wihlborg, A.K.; Niroula, A.; Swaminathan, B.; Johnsson, E.; Stephens, O.W.; Morgan, G.; Meissner, T.; Turesson, I.; et al. The multiple myeloma risk allele at 5q15 lowers ELL2 expression and increases ribosomal gene expression. Nat. Commun. 2018, 9, 1649. [Google Scholar] [CrossRef]
  377. Swaminathan, B.; Thorleifsson, G.; Jöud, M.; Ali, M.; Johnsson, E.; Ajore, R.; Sulem, P.; Halvarsson, B.M.; Eyjolfsson, G.; Haraldsdottir, V.; et al. Variants in ELL2 influencing immunoglobulin levels associate with multiple myeloma. Nat. Commun. 2015, 6, 12. [Google Scholar] [CrossRef]
  378. Martincic, K.; Alkan, S.A.; Cheatle, A.; Borghesi, L.; Milcarek, C. Transcription elongation factor ELL2 directs immunoglobulin secretion in plasma cells by stimulating altered RNA processing. Nat. Immunol. 2009, 10, 1102–1109. [Google Scholar] [CrossRef] [PubMed]
  379. Campa, D.; Martino, A.; Macauda, A.; Dudziński, M.; Suska, A.; Druzd-Sitek, A.; Raab, M.S.; Moreno, V.; Huhn, S.; Butrym, A.; et al. Genetic polymorphisms in genes of class switch recombination and multiple myeloma risk and survival: An IMMEnSE study. Leuk. Lymphoma 2019, 60, 1803–1811. [Google Scholar] [CrossRef]
  380. Roddam, P.L.; Rollinson, S.; O’Driscoll, M.; Jeggo, P.A.; Jack, A.; Morgan, G.J. Genetic variants of NHEJ DNA ligase IV can affect the risk of developing multiple myeloma, a tumour characterised by aberrant class switch recombination. J. Med. Genet. 2002, 39, 900–905. [Google Scholar] [CrossRef]
  381. Hayden, P.J.; Tewari, P.; Morris, D.W.; Staines, A.; Crowley, D.; Nieters, A.; Becker, N.; De sanjosé, S.; Foretova, L.; Maynadié, M.; et al. Variation in DNA repair genes XRCC3, XRCC4, XRCC5 and susceptibility to myeloma. Hum. Mol. Genet. 2007, 16, 3117–3127. [Google Scholar] [CrossRef]
  382. Elia, A.E.H.; Wang, D.C.; Willis, N.A.; Boardman, A.P.; Hajdu, I.; Adeyemi, R.O.; Lowry, E.; Gygi, S.P.; Scully, R.; Elledge, S.J. RFWD3-Dependent Ubiquitination of RPA Regulates Repair at Stalled Replication Forks. Mol. Cell 2015, 60, 280–293. [Google Scholar] [CrossRef] [PubMed]
  383. Inano, S.; Sato, K.; Katsuki, Y.; Kobayashi, W.; Tanaka, H.; Nakajima, K.; Nakada, S.; Miyoshi, H.; Knies, K.; Takaori-Kondo, A.; et al. RFWD3-Mediated Ubiquitination Promotes Timely Removal of Both RPA and RAD51 from DNA Damage Sites to Facilitate Homologous Recombination. Mol. Cell 2017, 66, 622–634.e8. [Google Scholar] [CrossRef] [PubMed]
  384. Duan, H.; Pathania, S. RPA, RFWD3 and BRCA2 at stalled forks: A balancing act. Mol. Cell. Oncol. 2020, 7, 1801089. [Google Scholar] [CrossRef]
  385. Gallina, I.; Hendriks, I.A.; Hoffmann, S.; Larsen, N.B.; Johansen, J.; Colding-Christensen, C.S.; Schubert, L.; Sellés-Baiget, S.; Fábián, Z.; Kühbacher, U.; et al. The ubiquitin ligase RFWD3 is required for translesion DNA synthesis. Mol. Cell 2021, 81, 442–458.e9. [Google Scholar] [CrossRef]
  386. Feeney, L.; Muñoz, I.M.; Lachaud, C.; Toth, R.; Appleton, P.L.; Schindler, D.; Rouse, J. RPA-Mediated Recruitment of the E3 Ligase RFWD3 Is Vital for Interstrand Crosslink Repair and Human Health. Mol. Cell 2017, 66, 610–621.e4. [Google Scholar] [CrossRef]
  387. Salzer, U.; Jennings, S.; Grimbacher, B. To switch or not to switch—The opposing roles of TACI in terminal B cell differentiation. Eur. J. Immunol. 2007, 37, 17–20. [Google Scholar] [CrossRef]
  388. Wang, T.; Liu, N.S.; Seet, L.F.; Hong, W. The emerging role of VHS domain-containing Tom1, Tom1L1 and Tom1L2 in membrane trafficking. Traffic 2010, 11, 1119–1128. [Google Scholar] [CrossRef]
  389. Pengo, N.; Scolari, M.; Oliva, L.; Milan, E.; Mainoldi, F.; Raimondi, A.; Fagioli, C.; Merlini, A.; Mariani, E.; Pasqualetto, E.; et al. Plasma cells require autophagy for sustainable immunoglobulin production. Nat. Immunol. 2013, 14, 298–305. [Google Scholar] [CrossRef] [PubMed]
  390. Cenci, S. Autophagy, a new determinant of plasma cell differentiation and antibody responses. Mol. Immunol. 2014, 62, 289–295. [Google Scholar] [CrossRef]
  391. Greenberg, A.J.; Lee, A.M.; Serie, D.J.; McDonnell, S.K.; Cerhan, J.R.; Liebow, M.; Larson, D.R.; Colby, C.L.; Norman, A.D.; Kyle, R.A.; et al. Single-nucleotide polymorphism rs1052501 associated with monoclonal gammopathy of undetermined significance and multiple myeloma. Leukemia 2013, 27, 515–516. [Google Scholar] [CrossRef]
  392. Preuss, F.; Chatterjee, D.; Mathea, S.; Shrestha, S.; St-Germain, J.; Saha, M.; Kannan, N.; Raught, B.; Rottapel, R.; Knapp, S. Nucleotide Binding, Evolutionary Insights, and Interaction Partners of the Pseudokinase Unc-51-like Kinase 4. Structure 2020, 28, 1184–1196.e6. [Google Scholar] [CrossRef] [PubMed]
  393. Kumar, A.; Rajendran, V.; Sethumadhavan, R.; Purohit, R. CEP proteins: The knights of centrosome dynasty. Protoplasma 2013, 250, 965–983. [Google Scholar] [CrossRef] [PubMed]
  394. Alexandrov, L.B.; Nik-Zainal, S.; Wedge, D.C.; Aparicio, S.A.J.R.; Behjati, S.; Biankin, A.V.; Bignell, G.R.; Bolli, N.; Borg, A.; Børresen-Dale, A.L.; et al. Signatures of mutational processes in human cancer. Nature 2013, 500, 415–421. [Google Scholar] [CrossRef] [PubMed]
  395. Hoang, P.H.; Cornish, A.J.; Dobbins, S.E.; Kaiser, M.; Houlston, R.S. Mutational processes contributing to the development of multiple myeloma. Blood Cancer J. 2019, 9, 60. [Google Scholar] [CrossRef]
  396. Lawrence, M.S.; Stojanov, P.; Polak, P.; Kryukov, G.V.; Cibulskis, K.; Sivachenko, A.; Carter, S.L.; Stewart, C.; Mermel, C.H.; Roberts, S.A.; et al. Mutational heterogeneity in cancer and the search for new cancer-associated genes. Nature 2013, 499, 214–218. [Google Scholar] [CrossRef] [PubMed]
  397. Liu, P.; Leong, T.; Quam, L.; Billadeau, D.; Kay, N.E.; Greipp, P.; Kyle, R.A.; Oken, M.M.; Van Ness, B. Activating Mutations of N- And K-ras in Multiple Myeloma Show Different Clinical Associations: Analysis of the Eastern Cooperative Oncology Group Phase III Trial. Blood 1996, 88, 2699–2706. [Google Scholar] [CrossRef]
  398. Hoang, P.H.; Dobbins, S.E.; Cornish, A.J.; Chubb, D.; Law, P.J.; Kaiser, M.; Houlston, R.S. Whole-genome sequencing of multiple myeloma reveals oncogenic pathways are targeted somatically through multiple mechanisms. Leukemia 2018, 32, 2459–2470. [Google Scholar] [CrossRef]
  399. Annunziata, C.M.; Davis, R.E.; Demchenko, Y.; Bellamy, W.; Gabrea, A.; Zhan, F.; Lenz, G.; Hanamura, I.; Wright, G.; Xiao, W.; et al. Frequent Engagement of the Classical and Alternative NF-κB Pathways by Diverse Genetic Abnormalities in Multiple Myeloma. Cancer Cell 2007, 12, 115–130. [Google Scholar] [CrossRef]
  400. Keats, J.J.; Fonseca, R.; Chesi, M.; Schop, R.; Baker, A.; Chng, W.J.; Van Wier, S.; Tiedemann, R.; Shi, C.X.; Sebag, M.; et al. Promiscuous Mutations Activate the Noncanonical NF-κB Pathway in Multiple Myeloma. Cancer Cell 2007, 12, 131–144. [Google Scholar] [CrossRef]
  401. Mithraprabhu, S.; Hocking, J.; Ramachandran, M.; Choi, K.; Klarica, D.; Khong, T.; Reynolds, J.; Spencer, A. DNA-repair gene mutations are highly prevalent in circulating tumour dna from multiple myeloma patients. Cancers 2019, 11, 917. [Google Scholar] [CrossRef]
  402. Jovanović, K.K.; Escure, G.; Demonchy, J.; Willaume, A.; Van De Wyngaert, Z.; Farhat, M.; Chauvet, P.; Facon, T.; Quesnel, B.; Manier, S. Deregulation and targeting of TP53 pathway in multiple myeloma. Front. Oncol. 2019, 9, 665. [Google Scholar] [CrossRef]
  403. Kunz, V.; Bommert, K.S.; Kruk, J.; Schwinning, D.; Chatterjee, M.; Stühmer, T.; Bargou, R.; Bommert, K. Targeting of the E3 ubiquitin-protein ligase HUWE1 impairs DNA repair capacity and tumor growth in preclinical multiple myeloma models. Sci. Rep. 2020, 10, 18419. [Google Scholar] [CrossRef]
  404. Dumontet, C.; Landi, S.; Reiman, T.; Perry, T.; Plesa, A.; Bellini, I.; Barale, R.; Pilarski, L.M.; Troncy, J.; Tavtigian, S.; et al. Genetic polymorphisms associated with outcome in multiple myeloma patients receiving high-dose melphalan. Bone Marrow Transplant. 2010, 45, 1316–1324. [Google Scholar] [CrossRef]
  405. Kumar, S.; Talluri, S.; Pal, J.; Yuan, X.; Lu, R.; Nanjappa, P.; Samur, M.K.; Munshi, N.C.; Shammas, M.A. Role of apurinic/apyrimidinic nucleases in the regulation of homologous recombination in myeloma: Mechanisms and translational significance. Blood Cancer J. 2018, 8, 92. [Google Scholar] [CrossRef]
  406. Keats, J.J.; Speyer, G.; Christofferson, A.; Legendre, C.; Aldrich, J.; Russell, M.; Cuyugan, L.; Adkins, J.; Blanski, A.; Hodges, M.; et al. Molecular Predictors of Outcome and Drug Response in Multiple Myeloma: An Interim Analysis of the Mmrf CoMMpass Study. Blood 2016, 128, 194. [Google Scholar] [CrossRef]
  407. Boyd, K.D.; Ross, F.M.; Walker, B.A.; Wardell, C.P.; Tapper, W.J.; Chiecchio, L.; Dagrada, G.P.; Konn, Z.J.; Gregory, W.M.; Jackson, G.H.; et al. Mapping of chromosome 1p deletions in myeloma identifies FAM46C at 1p12 and CDKN2C at 1p32.3 as being genes in regions associated with adverse survival. Clin. Cancer Res. 2011, 17, 7776–7784. [Google Scholar] [CrossRef]
  408. Barbieri, M.; Manzoni, M.; Fabris, S.; Ciceri, G.; Todoerti, K.; Simeon, V.; Musto, P.; Cortelezzi, A.; Baldini, L.; Neri, A.; et al. Compendium of FAM46C gene mutations in plasma cell dyscrasias. Br. J. Haematol. 2016, 174, 642–645. [Google Scholar] [CrossRef]
  409. Mroczek, S.; Chlebowska, J.; Kuliński, T.M.; Gewartowska, O.; Gruchota, J.; Cysewski, D.; Liudkovska, V.; Borsuk, E.; Nowis, D.; Dziembowski, A. The non-canonical poly(A) polymerase FAM46C acts as an onco-suppressor in multiple myeloma. Nat. Commun. 2017, 8, 619. [Google Scholar] [CrossRef] [PubMed]
  410. Zhang, H.; Zhang, S.; Hu, J.; Wu, Y.; Ma, X.; Chen, Y.; Yu, B.; Liao, S.; Huang, H.; Gao, S. Structural and functional characterization of multiple myeloma associated cytoplasmic poly(A) polymerase FAM46C. Cancer Commun. 2021, 41, 615. [Google Scholar] [CrossRef] [PubMed]
  411. Herrero, A.B.; Quwaider, D.; Corchete, L.A.; Mateos, M.V.; García-Sanz, R.; Gutiérrez, N.C. FAM46C controls antibody production by the polyadenylation of immunoglobulin mRNAs and inhibits cell migration in multiple myeloma. J. Cell. Mol. Med. 2020, 24, 4171–4182. [Google Scholar] [CrossRef] [PubMed]
  412. Kuchta, K.; Muszewska, A.; Knizewski, L.; Steczkiewicz, K.; Wyrwicz, L.S.; Pawlowski, K.; Rychlewski, L.; Ginalski, K. FAM46 proteins are novel eukaryotic non-canonical poly(A) polymerases. Nucleic Acids Res. 2016, 44, 3534–3548. [Google Scholar] [CrossRef] [PubMed]
  413. Zhou, Z.; Gong, Q.; Wang, Y.; Li, M.; Wang, L.; Ding, H.; Li, P. The biological function and clinical significance of SF3B1 mutations in cancer. Biomark. Res. 2020, 8, 38. [Google Scholar] [CrossRef]
  414. Hofman, I.J.F.; Patchett, S.; Van Duin, M.; Geerdens, E.; Verbeeck, J.; Michaux, L.; Delforge, M.; Sonneveld, P.; Johnson, A.W.; De Keersmaecker, K. Low frequency mutations in ribosomal proteins RPL10 and RPL5 in multiple myeloma. Haematologica 2017, 102, e317–e320. [Google Scholar] [CrossRef] [PubMed]
  415. Pawlyn, C.; Kaiser, M.F.; Heuck, C.; Melchor, L.; Wardell, C.P.; Murison, A.; Chavan, S.S.; Johnson, D.C.; Begum, D.B.; Dahir, N.M.; et al. The spectrum and clinical impact of epigenetic modifier mutations in Myeloma. Clin. Cancer Res. 2016, 22, 5783–5794. [Google Scholar] [CrossRef]
  416. Fyodorov, D.V.; Zhou, B.R.; Skoultchi, A.I.; Bai, Y. Emerging roles of linker histones in regulating chromatin structure and function. Nat. Rev. Mol. Cell Biol. 2018, 19, 192–206. [Google Scholar] [CrossRef] [PubMed]
  417. Dupéré-Richer, D.; Licht, J.D. Epigenetic regulatory mutations and epigenetic therapy for multiple myeloma. Curr. Opin. Hematol. 2017, 24, 336–344. [Google Scholar] [CrossRef] [PubMed]
  418. Martinez-Garcia, E.; Popovic, R.; Min, D.J.; Sweet, S.M.M.; Thomas, P.M.; Zamdborg, L.; Heffner, A.; Will, C.; Lamy, L.; Staudt, L.M.; et al. The MMSET histone methyl transferase switches global histone methylation and alters gene expression in t(4;14) multiple myeloma cells. Blood 2011, 117, 211–220. [Google Scholar] [CrossRef]
  419. Popovic, R.; Martinez-Garcia, E.; Giannopoulou, E.G.; Zhang, Q.; Zhang, Q.; Ezponda, T.; Shah, M.Y.; Zheng, Y.; Will, C.M.; Small, E.C.; et al. Histone Methyltransferase MMSET/NSD2 Alters EZH2 Binding and Reprograms the Myeloma Epigenome through Global and Focal Changes in H3K36 and H3K27 Methylation. PLoS Genet. 2014, 10, e1004566. [Google Scholar] [CrossRef]
  420. Van Haaften, G.; Dalgliesh, G.L.; Davies, H.; Chen, L.; Bignell, G.; Greenman, C.; Edkins, S.; Hardy, C.; O’Meara, S.; Teague, J.; et al. Somatic mutations of the histone H3K27 demethylase gene UTX in human cancer. Nat. Genet. 2009, 41, 521–523. [Google Scholar] [CrossRef]
  421. Attar, N.; Kurdistani, S.K. Exploitation of EP300 and CREBBP lysine acetyltransferases by cancer. Cold Spring Harb. Perspect. Med. 2017, 7, a026534. [Google Scholar] [CrossRef]
  422. Heuser, M.; Yun, H.; Thol, F. Epigenetics in myelodysplastic syndromes. Semin. Cancer Biol. 2018, 51, 170–179. [Google Scholar] [CrossRef] [PubMed]
  423. Conery, A.R.; Centore, R.C.; Neiss, A.; Keller, P.J.; Joshi, S.; Spillane, K.L.; Sandy, P.; Hatton, C.; Pardo, E.; Zawadzke, L.; et al. Bromodomain inhibition of the transcriptional coactivators CBP/EP300 as a therapeutic strategy to target the IRF4 network in multiple myeloma. Elife 2016, 5, e10483. [Google Scholar] [CrossRef] [PubMed]
  424. Yamamoto, J.; Suwa, T.; Murase, Y.; Tateno, S.; Mizutome, H.; Asatsuma-Okumura, T.; Shimizu, N.; Kishi, T.; Momose, S.; Kizaki, M.; et al. ARID2 is a pomalidomide-dependent CRL4CRBN substrate in multiple myeloma cells. Nat. Chem. Biol. 2020, 16, 1208–1217. [Google Scholar] [CrossRef] [PubMed]
  425. Walker, B.A.; Wardell, C.P.; Chiecchio, L.; Smith, E.M.; Boyd, K.D.; Neri, A.; Davies, F.E.; Ross, F.M.; Morgan, G.J. Aberrant global methylation patterns affect the molecular pathogenesis and prognosis of multiple myeloma. Blood 2011, 117, 553–562. [Google Scholar] [CrossRef] [PubMed]
  426. Salhia, B.; Baker, A.; Ahmann, G.; Auclair, D.; Fonseca, R.; Carpten, J.D. DNA methylation analysis determines the high frequency of genic hypomethylation and low frequency of hypermethylation events in plasma cell tumors. Cancer Res. 2010, 70, 6934–6944. [Google Scholar] [CrossRef] [PubMed]
  427. Heuck, C.J.; Mehta, J.; Bhagat, T.; Gundabolu, K.; Yu, Y.; Khan, S.; Chrysofakis, G.; Schinke, C.; Tariman, J.; Vickrey, E.; et al. Myeloma Is Characterized by Stage-Specific Alterations in DNA Methylation That Occur Early during Myelomagenesis. J. Immunol. 2013, 190, 2966–2975. [Google Scholar] [CrossRef]
  428. Agirre, X.; Castellano, G.; Pascual, M.; Heath, S.; Kulis, M.; Segura, V.; Bergmann, A.; Esteve, A.; Merkel, A.; Raineri, E.; et al. Whole-epigenome analysis in multiple myeloma reveals DNA hypermethylation of B cell-specific enhancers. Genome Res. 2015, 25, 478–487. [Google Scholar] [CrossRef]
  429. Choudhury, S.R.; Ashby, C.; Tytarenko, R.; Bauer, M.; Wang, Y.; Deshpande, S.; Den, J.; Schinke, C.; Zangari, M.; Thanendrarajan, S.; et al. The functional epigenetic landscape of aberrant gene expression in molecular subgroups of newly diagnosed multiple myeloma. J. Hematol. Oncol. 2020, 13, 108. [Google Scholar] [CrossRef]
  430. Alberge, J.B.; Magrangeas, F.; Wagner, M.; Denié, S.; Guérin-Charbonnel, C.; Campion, L.; Attal, M.; Avet-Loiseau, H.; Carell, T.; Moreau, P.; et al. DNA hydroxymethylation is associated with disease severity and persists at enhancers of oncogenic regions in multiple myeloma. Clin. Epigenetics 2020, 12, 163. [Google Scholar] [CrossRef]
  431. Mullighan, C.G. TET2 mutations in myelodysplasia and myeloid malignancies. Nat. Genet. 2009, 41, 766–767. [Google Scholar] [CrossRef]
  432. Duployez, N.; Goursaud, L.; Fenwarth, L.; Bories, C.; Marceau-Renaut, A.; Boyer, T.; Fournier, E.; Nibourel, O.; Roche-Lestienne, C.; Huet, G.; et al. Familial myeloid malignancies with germline TET2 mutation. Leukemia 2020, 34, 1450–1453. [Google Scholar] [CrossRef]
  433. Solary, E.; Bernard, O.A.; Tefferi, A.; Fuks, F.; Vainchenker, W. The Ten-Eleven Translocation-2 (TET2) gene in hematopoiesis and hematopoietic diseases. Leukemia 2014, 28, 485–496. [Google Scholar] [CrossRef] [PubMed]
  434. Lio, C.W.J.; Shukla, V.; Samaniego-Castruita, D.; González-Avalos, E.; Chakraborty, A.; Yue, X.; Schatz, D.G.; Ay, F.; Rao, A. TET enzymes augment activation-induced deaminase (AID) expression via 5-hydroxymethylcytosine modifications at the Aicda superenhancer. Sci. Immunol. 2019, 4, eaau7523. [Google Scholar] [CrossRef]
  435. Rogozin, I.B.; Pavlov, Y.I.; Bebenek, K.; Matsuda, T.; Kunkel, T.A. Somatic mutation hotspots correlate with DNA polymerase eta error spectrum. Nat. Immunol. 2001, 2, 530–536. [Google Scholar] [CrossRef] [PubMed]
  436. Zhao, S.; Choi, M.; Heuck, C.; Mane, S.; Barlogie, B.; Lifton, R.P.; Dhodapkar, M. V Serial exome analysis of disease progression in premalignant gammopathies. Leukemia 2014, 28, 1548–1552. [Google Scholar] [CrossRef] [PubMed]
  437. Bustoros, M.; Park, J.; Salem, K.Z.; Liu, C.-J.; Capelletti, M.; Huynh, D.; Tai, Y.-T.; Mouhieddine, T.H.; Freeman, S.; Ha, G.; et al. Next Generation Sequencing Identifies Smoldering Multiple Myeloma Patients with a High Risk of Disease Progression. Blood 2017, 130, 392. [Google Scholar]
  438. Maura, F.; Petljak, M.; Lionetti, M.; Cifola, I.; Liang, W.; Pinatel, E.; Alexandrov, L.B.; Fullam, A.; Martincorena, I.; Dawson, K.J.; et al. Biological and prognostic impact of APOBEC-induced mutations in the spectrum of plasma cell dyscrasias and multiple myeloma cell lines. Leukemia 2018, 32, 1043–1047. [Google Scholar] [CrossRef]
  439. Manzoni, M.; Marchica, V.; Storti, P.; Ziccheddu, B.; Sammarelli, G.; Todaro, G.; Pelizzoni, F.; Salerio, S.; Notarfranchi, L.; Pompa, A.; et al. Application of next-generation sequencing for the genomic characterization of patients with smoldering myeloma. Cancers 2020, 12, 1332. [Google Scholar] [CrossRef]
  440. Egan, J.B.; Kortuem, K.M.; Kurdoglu, A.; Izatt, T.; Aldrich, J.; Reiman, R.; Phillips, L.; Baker, A.; Shi, C.X.; Schmidt, J.; et al. Extramedullary myeloma whole genome sequencing reveals novel mutations in Cereblon, proteasome subunit G2 and the glucocorticoid receptor in multi drug resistant disease. Br. J. Haematol. 2013, 161, 748–751. [Google Scholar] [CrossRef]
  441. Kortüm, K.M.; Mai, E.K.; Hanafiah, N.H.; Shi, C.X.; Zhu, Y.X.; Bruins, L.; Barrio, S.; Jedlowski, P.; Merz, M.; Xu, J.; et al. Targeted sequencing of refractory myeloma reveals a high incidence of mutations in CRBN and Ras pathway genes. Blood 2016, 128, 1226–1233. [Google Scholar] [CrossRef]
  442. Barrio, S.; Stühmer, T.; Da-Viá, M.; Barrio-Garcia, C.; Lehners, N.; Besse, A.; Cuenca, I.; Garitano-Trojaola, A.; Fink, S.; Leich, E.; et al. Spectrum and functional validation of PSMB5 mutations in multiple myeloma. Leukemia 2019, 33, 447–456. [Google Scholar] [CrossRef]
  443. Gooding, S.; Ansari-Pour, N.; Towfic, F.; Estévez, M.O.; Chamberlain, P.P.; Tsai, K.T.; Flynt, E.; Hirst, M.; Rozelle, D.; Dhiman, P.; et al. Multiple cereblon genetic changes are associated with acquired resistance to lenalidomide or pomalidomide in multiple myeloma. Blood 2021, 137, 232–237. [Google Scholar] [CrossRef]
  444. Cohen, Y.C.; Zada, M.; Wang, S.Y.; Bornstein, C.; David, E.; Moshe, A.; Li, B.; Shlomi-Loubaton, S.; Gatt, M.E.; Gur, C.; et al. Identification of resistance pathways and therapeutic targets in relapsed multiple myeloma patients through single-cell sequencing. Nat. Med. 2021, 27, 491–503. [Google Scholar] [CrossRef]
  445. Iskierka-Jażdżewska, E.; Canzian, F.; Stępień, A.; Martino, A.; Campa, D.; Stein, A.; Krawczyk-Kuliś, M.; Rybicka-Ramos, M.; Kyrcz-Krzemień, S.; Butrym, A.; et al. Cereblon (CRBN) gene polymorphisms predict clinical response and progression-free survival in relapsed/refractory multiple myeloma patients treated with lenalidomide: A pharmacogenetic study from the IMMEnSE consortium. Leuk. Lymphoma 2020, 61, 699–706. [Google Scholar] [CrossRef]
  446. Haertle, L.; Barrio, S.; Munawar, U.; Han, S.; Zhou, X.; Vogt, C.; Alonso, R.; Bittrich, M.; Ruiz-Heredia, Y.; Da-Via, M.; et al. Cereblon Enhancer Methylation and IMiD Resistance in Multiple Myeloma. Blood 2021, in press. [Google Scholar] [CrossRef] [PubMed]
  447. Rogozin, I.B.; Pavlov, Y.I.; Goncearenco, A.; De, S.; Lada, A.G.; Poliakov, E.; Panchenko, A.R.; Cooper, D.N. Mutational signatures and mutable motifs in cancer genomes. Brief. Bioinform. 2017, 19, 1085–1101. [Google Scholar] [CrossRef] [PubMed]
  448. Alexandrov, L.B.; Kim, J.; Haradhvala, N.J.; Huang, M.N.; Tian Ng, A.W.; Wu, Y.; Boot, A.; Covington, K.R.; Gordenin, D.A.; Bergstrom, E.N.; et al. The repertoire of mutational signatures in human cancer. Nature 2020, 578, 94–101. [Google Scholar] [CrossRef]
  449. Roberts, S.A.; Gordenin, D.A. Hypermutation in human cancer genomes: Footprints and mechanisms. Nat. Rev. Cancer 2014, 14, 786–800. [Google Scholar] [CrossRef]
  450. Conticello, S.G. The AID/APOBEC family of nucleic acid mutators. Genome Biol. 2008, 9, 229. [Google Scholar] [CrossRef] [PubMed]
  451. Pham, P.; Bransteitter, R.; Petruska, J.; Goodman, M.F. Processive AID-catalysed cytosine deamination on single-stranded DNA simulates somatic hypermutation. Nature 2003, 424, 103–107. [Google Scholar] [CrossRef]
  452. Nik-Zainal, S.; Alexandrov, L.B.; Wedge, D.C.; Van Loo, P.; Greenman, C.D.; Raine, K.; Jones, D.; Hinton, J.; Marshall, J.; Stebbings, L.A.; et al. Mutational processes molding the genomes of 21 breast cancers. Cell 2012, 149, 979–993. [Google Scholar] [CrossRef]
  453. Lada, A.G.; Dhar, A.; Boissy, R.J.; Hirano, M.; Rubel, A.A.; Rogozin, I.B.; Pavlov, Y.I. AID/APOBEC cytosine deaminase induces genome-wide kataegis. Biol. Direct 2012, 7, 47. [Google Scholar] [CrossRef] [PubMed]
  454. Saini, N.; Gordenin, D.A. Hypermutation in single-stranded DNA. DNA Repair (Amst) 2020, 91–92, 102868. [Google Scholar] [CrossRef]
  455. Mas-Ponte, D.; Supek, F. DNA mismatch repair promotes APOBEC3-mediated diffuse hypermutation in human cancers. Nat. Genet. 2020, 52, 958–968. [Google Scholar] [CrossRef]
  456. Walker, B.A.; Wardell, C.P.; Murison, A.; Boyle, E.M.; Begum, D.B.; Dahir, N.M.; Proszek, P.Z.; Melchor, L.; Pawlyn, C.; Kaiser, M.F.; et al. APOBEC family mutational signatures are associated with poor prognosis translocations in multiple myeloma. Nat. Commun. 2015, 6, 6997. [Google Scholar] [CrossRef]
  457. Maura, F.; Degasperi, A.; Nadeu, F.; Leongamornlert, D.; Davies, H.; Moore, L.; Royo, R.; Ziccheddu, B.; Puente, X.S.; Avet-Loiseau, H.; et al. A practical guide for mutational signature analysis in hematological malignancies. Nat. Commun. 2019, 10, 2969. [Google Scholar] [CrossRef]
  458. Maura, F.; Landgren, O.; Morgan, G.J. Designing Evolutionary-based Interception Strategies to Block the Transition from Precursor Phases to Multiple Myeloma. Clin. Cancer Res. 2021, 27, 15–23. [Google Scholar] [CrossRef] [PubMed]
  459. Roberts, S.A.; Sterling, J.; Thompson, C.; Harris, S.; Mav, D.; Shah, R.; Klimczak, L.J.; Kryukov, G.V.; Malc, E.; Mieczkowski, P.A.; et al. Clustered Mutations in Yeast and in Human Cancers Can Arise from Damaged Long Single-Strand DNA Regions. Mol. Cell 2012, 46, 424–435. [Google Scholar] [CrossRef]
  460. Rogozin, I.B.; Kolchanov, N.A. Somatic hypermutagenesis in immunoglobulin genes. II. Influence of neighbouring base sequences on mutagenesis. BBA-Gene Struct. Expr. 1992, 1171, 11–18. [Google Scholar] [CrossRef]
  461. Pavlov, Y.I.; Rogozin, I.B.; Galkin, A.P.; Aksenova, A.Y.; Hanaoka, F.; Rada, C.; Kunkel, T.A. Correlation of somatic hypermutation specificity and A-T base pair substitution errors by DNA polymerase eta during copying of a mouse immunoglobulin kappa light chain transgene. Proc. Natl. Acad. Sci. USA 2002, 99, 9954–9959. [Google Scholar] [CrossRef] [PubMed]
  462. Rogozin, I.B.; Roche-Lima, A.; Tyryshkin, K.; Carrasquillo-Carrión, K.; Lada, A.G.; Poliakov, L.Y.; Schwartz, E.; Saura, A.; Yurchenko, V.; Cooper, D.N.; et al. DNA Methylation, Deamination, and Translesion Synthesis Combine to Generate Footprint Mutations in Cancer Driver Genes in B-Cell Derived Lymphomas and Other Cancers. Front. Genet. 2021, 12, 783. [Google Scholar] [CrossRef]
  463. Taylor, B.J.M.; Nik-Zainal, S.; Wu, Y.L.; Stebbings, L.A.; Raine, K.; Campbell, P.J.; Rada, C.; Stratton, M.R.; Neuberger, M.S. DNA deaminases induce break-associated mutation showers with implication of APOBEC3B and 3A in breast cancer kataegis. Elife 2013, 2, e00534. [Google Scholar] [CrossRef]
  464. Chan, K.; Roberts, S.A.; Klimczak, L.J.; Sterling, J.F.; Saini, N.; Malc, E.P.; Kim, J.; Kwiatkowski, D.J.; Fargo, D.C.; Mieczkowski, P.A.; et al. An APOBEC3A hypermutation signature is distinguishable from the signature of background mutagenesis by APOBEC3B in human cancers. Nat. Genet. 2015, 47, 1067–1072. [Google Scholar] [CrossRef]
  465. Rogozin, I.; Roche-Lima, A.; Lada, A.; Belinky, F.; Sidorenko, I.; Glazko, G.; Babenko, V.; Cooper, D.; Pavlov, Y. Nucleotide Weight Matrices Reveal Ubiquitous Mutational Footprints of AID/APOBEC Deaminases in Human Cancer Genomes. Cancers 2019, 11, 211. [Google Scholar] [CrossRef]
  466. Álvarez-Prado, Á.F.; Pérez-Durán, P.; Pérez-García, A.; Benguria, A.; Torroja, C.; de Yébenes, V.G.; Ramiro, A.R. A broad atlas of somatic hypermutation allows prediction of activation-induced deaminase targets. J. Exp. Med. 2018, 215, 761–771. [Google Scholar] [CrossRef]
  467. Rogozin, I.B.; Lada, A.G.; Goncearenco, A.; Green, M.R.; De, S.; Nudelman, G.; Panchenko, A.R.; Koonin, E.V.; Pavlov, Y.I. Activation induced deaminase mutational signature overlaps with CpG methylation sites in follicular lymphoma and other cancers. Sci. Rep. 2016, 6, 38133. [Google Scholar] [CrossRef] [PubMed]
  468. Maura, F.; Rustad, E.H.; Yellapantula, V.; Łuksza, M.; Hoyos, D.; Maclachlan, K.H.; Diamond, B.T.; Greenbaum, B.D.; Morgan, G.; Lesokhin, A.; et al. Role of AID in the temporal pattern of acquisition of driver mutations in multiple myeloma. Leukemia 2020, 34, 1476–1480. [Google Scholar] [CrossRef] [PubMed]
  469. Maul, R.W.; Gearhart, P.J. AID and somatic hypermutation. In Advances in Immunology; Academic Press Inc.: AEAmsterdam, The Netherlands, 2010; Volume 105, pp. 159–191. [Google Scholar]
  470. Ross, F.M.; Chiecchio, L.; Dagrada, G.; Protheroe, R.K.M.; Stockley, D.M.; Harrison, C.J.; Cross, N.C.; Szubert, A.J.; Drayson, M.T.; Morgan, G.J. The t(14;20) is a poor prognostic factor in myeloma but is associated with long-term stable disease in monoclonal gammopathies of undetermined significance. Haematologica 2010, 95, 1221–1225. [Google Scholar] [CrossRef] [PubMed]
  471. Rouf Banday, A.; Onabajo, O.O.; Lin, S.H.Y.; Obajemu, A.; Vargas, J.M.; Delviks-Frankenberry, K.A.; Lamy, P.; Bayanjargal, A.; Zettelmeyer, C.; Florez-Vargas, O.; et al. Targeting natural splicing plasticity of APOBEC3B restricts its expression and mutagenic activity. Commun. Biol. 2021, 4, 386. [Google Scholar] [CrossRef]
  472. Flactif, M.; Zandecki, M.; Laï, J.L.; Bernardi, F.; Obein, V.; Bauters, F.; Facon, T. Interphase fluorescence in situ hybridization (FISH) as a powerful tool for the detection of aneuploidy in multiple myeloma. Leukemia 1995, 9, 2109–2114. [Google Scholar]
  473. Drach, J.; Schuster, J.; Nowotny, H.; Angerler, J.; Rosenthal, F.; Fiegl, M.; Rothermundt, C.; Gsur, A.; Jager, U.; Heinz, R.; et al. Multiple myeloma: High incidence of chromosomal aneuploidy as detected by interphase fluorescence in situ hybridization. Cancer Res. 1995, 55, 3854–3859. [Google Scholar]
  474. Lee, W.; Han, K.; Drut, R.M.; Harris, C.P.; Meisner, L.F. Use of fluorescence in situ hybridization for retrospective detection of aneuploidy in multiple myeloma. Genes Chromosom. Cancer 1993, 7, 137–143. [Google Scholar] [CrossRef]
  475. Pratt, G. Molecular aspects of multiple myeloma. J. Clin. Pathol.-Mol. Pathol. 2002, 55, 273–283. [Google Scholar] [CrossRef]
  476. Walker, B.A.; Leone, P.E.; Chiecchio, L.; Dickens, N.J.; Jenner, M.W.; Boyd, K.D.; Johnson, D.C.; Gonzalez, D.; Dagrada, G.P.; Protheroe, R.K.M.; et al. A compendium of myeloma-associated chromosomal copy number abnormalities and their prognostic value. Blood 2010, 116, e56–e65. [Google Scholar] [CrossRef] [PubMed]
  477. Chretien, M.L.; Corre, J.; Lauwers-Cances, V.; Magrangeas, F.; Cleynen, A.; Yon, E.; Hulin, C.; Leleu, X.; Orsini-Piocelle, F.; Blade, J.S.; et al. Understanding the role of hyperdiploidy in myeloma prognosis: Which trisomies really matter? Blood 2015, 126, 2713–2719. [Google Scholar] [CrossRef]
  478. Aktas Samur, A.; Minvielle, S.; Shammas, M.; Fulciniti, M.; Magrangeas, F.; Richardson, P.G.; Moreau, P.; Attal, M.; Anderson, K.C.; Parmigiani, G.; et al. Deciphering the chronology of copy number alterations in Multiple Myeloma. Blood Cancer J. 2019, 9, 39. [Google Scholar] [CrossRef]
  479. Barilà, G.; Bonaldi, L.; Grassi, A.; Martines, A.; Liço, A.; Macrì, N.; Nalio, S.; Pavan, L.; Berno, T.; Branca, A.; et al. Identification of the true hyperdiploid multiple myeloma subset by combining conventional karyotyping and FISH analysis. Blood Cancer J. 2020, 10, 18. [Google Scholar] [CrossRef] [PubMed]
  480. Rasillo, A.; Tabernero, M.D.; Sánchez, M.L.; Pérez, D.A.M.; Martín, A.M.; Hernández, J.; Moro, M.J.; Fernández-Calvo, J.; Sayagués, J.M.; Bortoluci, A.; et al. Fluorescence in situ hybridization analysis of aneuploidization patterns in monoclonal gammopathy of undetermined significance versus multiple myeloma and plasma cell leukemia. Cancer 2003, 97, 601–609. [Google Scholar] [PubMed]
  481. Brousseau, M.; Leleu, X.; Gerard, J.; Gastinne, T.; Godon, A.; Genevieve, F.; Dib, M.; Lai, J.L.; Facon, T.; Zandecki, M. Hyperdiploidy is a common finding in monoclonal gammopathy of undetermined significance and monosomy 13 is restricted to these hyperdiploid patients. Clin. Cancer Res. 2007, 13, 6026–6031. [Google Scholar] [CrossRef] [PubMed]
  482. Fonseca, R.; Barlogie, B.; Bataille, R.; Bastard, C.; Bergsagel, P.L.; Chesi, M.; Davies, F.E.; Drach, J.; Greipp, P.R.; Kirsch, I.R.; et al. Genetics and Cytogenetics of Multiple Myeloma: A Workshop Report. Cancer Res. 2004, 64, 1546–1558. [Google Scholar] [CrossRef]
  483. Bergsagel, P.L.; Chesi, M.; Nardini, E.; Brents, L.A.; Kirby, S.L.; Kuehl, W.M. Promiscuous translocations into immunoglobulin heavy chain switch regions in multiple myeloma. Proc. Natl. Acad. Sci. USA 1996, 93, 13931–13936. [Google Scholar] [CrossRef] [PubMed]
  484. Leif Bergsagel, P.; Michael Kuehl, W. Chromosome translocations in multiple myeloma. Oncogene 2001, 20, 5611–5622. [Google Scholar] [CrossRef] [PubMed]
  485. Chesi, M.; Nardini, E.; Lim, R.S.C.; Smith, K.D.; Michael Kuehl, W.; Bergsagel, P.L. The t(4;14) translocation in myeloma dysregulates both FGFR3 and a novel gene, MMSET, resulting in IgH/MMSET hybrid transcripts. Blood 1998, 92, 3025–3034. [Google Scholar] [CrossRef] [PubMed]
  486. Kalff, A.; Spencer, A. The t(4;14) translocation and FGFR3 overexpression in multiple myeloma: Prognostic implications and current clinical strategies. Blood Cancer J. 2012, 2, e89. [Google Scholar] [CrossRef]
  487. Plowright, E.E.; Li, Z.; Bergsagel, P.L.; Chesi, M.; Barber, D.L.; Branch, D.R.; Hawley, R.G.; Stewart, A.K. Ectopic expression of fibroblast growth factor receptor 3 promotes myeloma cell proliferation and prevents apoptosis. Blood 2000, 95, 992–998. [Google Scholar] [CrossRef]
  488. Alvarez, J.V.; Frank, D.A. Genome-wide analysis of STAT target genes: Elucidating the mechanism of STAT-mediated oncogenesis. Cancer Biol. Ther. 2004, 3, 1045–1050. [Google Scholar] [CrossRef]
  489. Ramlee, M.K.; Wang, J.; Toh, W.X.; Li, S. Transcription regulation of the human telomerase reverse transcriptase (hTERT) gene. Genes 2016, 7, 50. [Google Scholar] [CrossRef]
  490. Bharti, A.C.; Shishodia, S.; Reuben, J.M.; Weber, D.; Alexanian, R.; Raj-Vadhan, S.; Estrov, Z.; Talpaz, M.; Aggarwal, B.B. Nuclear factor-κB and STAT3 are constitutively active in CD138 + cells derived from multiple myeloma patients, and suppression of these transcription factors leads to apoptosis. Blood 2004, 103, 3175–3184. [Google Scholar] [CrossRef]
  491. Catlett-Falcone, R.; Landowski, T.H.; Oshiro, M.M.; Turkson, J.; Levitzki, A.; Savino, R.; Ciliberto, G.; Moscinski, L.; Fernández-Luna, J.L.; Nuñez, G.; et al. Constitutive activation of Stat3 signaling confers resistance to apoptosis in human U266 myeloma cells. Immunity 1999, 10, 105–115. [Google Scholar] [CrossRef]
  492. Brocke-Heidrich, K.; Kretzschmar, A.K.; Pfeifer, G.; Henze, C.; Löffler, D.; Koczan, D.; Thiesen, H.J.; Burger, R.; Gramatzki, M.; Horn, F. Interleukin-6-dependent gene expression profiles in multiple myeloma INA-6 cells reveal a Bcl-2 family-independent survival pathway closely associated with Stat3 activation. Blood 2004, 103, 242–251. [Google Scholar] [CrossRef]
  493. Avery, D.T.; Deenick, E.K.; Ma, C.S.; Suryani, S.; Simpson, N.; Chew, G.Y.; Chan, T.D.; Palendira, U.; Bustamante, J.; Boisson-Dupuis, S.; et al. B cell-intrinsic signaling through IL-21 receptor and STAT3 is required for establishing long-lived antibody responses in humans. J. Exp. Med. 2010, 207, 155–171. [Google Scholar] [CrossRef] [PubMed]
  494. Faris, M.; Kokot, N.; Stahl, N.; Nel, A.E. Involvement of Stat3 in interleukin-6-induced IgM production in a human B-cell line. Immunology 1997, 90, 350–357. [Google Scholar] [CrossRef] [PubMed]
  495. Ronchetti, D.; Greco, A.; Compasso, S.; Colombo, G.; Dell’Era, P.; Otsuki, T.; Lombardi, L.; Neri, A. Deregulated FGFR3 mutants in multiple myeloma cell lines with t(4;14): Comparative analysis of Y373C, K650E and the novel G384D mutations. Oncogene 2001, 20, 3553–3562. [Google Scholar] [CrossRef] [PubMed]
  496. Kuo, A.J.; Cheung, P.; Chen, K.; Zee, B.M.; Kioi, M.; Lauring, J.; Xi, Y.; Park, B.H.; Shi, X.; Garcia, B.A.; et al. NSD2 Links Dimethylation of Histone H3 at Lysine 36 to Oncogenic Programming. Mol. Cell 2011, 44, 609–620. [Google Scholar] [CrossRef]
  497. Shah, M.Y.; Martinez-Garcia, E.; Phillip, J.M.; Chambliss, A.B.; Popovic, R.; Ezponda, T.; Small, E.C.; Will, C.; Phillip, M.P.; Neri, P.; et al. MMSET/WHSC1 enhances DNA damage repair leading to an increase in resistance to chemotherapeutic agents. Oncogene 2016, 35, 5905–5915. [Google Scholar] [CrossRef]
  498. Park, J.W.; Chae, Y.C.; Kim, J.Y.; Oh, H.; Seo, S.B. Methylation of Aurora kinase A by MMSET reduces p53 stability and regulates cell proliferation and apoptosis. Oncogene 2018, 37, 6212–6224. [Google Scholar] [CrossRef]
  499. Pei, H.; Zhang, L.; Luo, K.; Qin, Y.; Chesi, M.; Fei, F.; Bergsagel, P.L.; Wang, L.; You, Z.; Lou, Z. MMSET regulates histone H4K20 methylation and 53BP1 accumulation at DNA damage sites. Nature 2011, 470, 124–129. [Google Scholar] [CrossRef]
  500. Evans, D.L.; Zhang, H.; Ham, H.; Pei, H.; Lee, S.B.; Kim, J.J.; Billadeau, D.D.; Lou, Z. MMSET is dynamically regulated during cell-cycle progression and promotes normal DNA replication. Cell Cycle 2016, 15, 95–105. [Google Scholar] [CrossRef]
  501. de Krijger, I.; van der Torre, J.; Peuscher, M.H.; Eder, M.; Jacobs, J.J.L. H3K36 dimethylation by MMSET promotes classical non-homologous end-joining at unprotected telomeres. Oncogene 2020, 39, 4814–4827. [Google Scholar] [CrossRef]
  502. Nguyen, H.V.; Dong, J.; Panchakshari, R.A.; Kumar, V.; Alt, F.W.; Bories, J.C. Histone methyltransferase MMSET promotes AID-mediated DNA breaks at the donor switch region during class switch recombination. Proc. Natl. Acad. Sci. USA 2017, 114, E10560–E10567. [Google Scholar] [CrossRef]
  503. Gao, G.; Smith, D.I. WWOX, large common fragile site genes, and cancer. Exp. Biol. Med. 2015, 240, 285–295. [Google Scholar] [CrossRef]
  504. Mangelsdorf, M.; Ried, K.; Woollatt, E.; Dayan, S.; Eyre, H.; Finnis, M.; Hobson, L.; Nancarrow, J.; Venter, D.; Baker, E.; et al. Chromosomal fragile site FRA16D and DNA instability in cancer. Cancer Res. 2000, 60, 1683–1689. [Google Scholar]
  505. Boersma-Vreugdenhil, G.R.; Kuipers, J.; Van Stralen, E.; Peeters, T.; Michaux, L.; Hagemeijer, A.; Pearson, P.L.; Clevers, H.C.; Bast, B.J.E.G. The recurrent translocation t(14;20)(q32;q12) in multiple myeloma results in aberrant expression of MAFB: A molecular and genetic analysis of the chromosomal breakpoint. Br. J. Haematol. 2004, 126, 355–363. [Google Scholar] [CrossRef]
  506. Chesi, M.; Bergsagel, P.L. Molecular pathogenesis of multiple myeloma: Basic and clinical updates. Int. J. Hematol. 2013, 97, 313–323. [Google Scholar] [CrossRef]
  507. Bergsagel, P.L.; Kuehl, W.M. Molecular pathogenesis and a consequent classification of multiple myeloma. J. Clin. Oncol. 2005, 23, 6333–6338. [Google Scholar] [CrossRef] [PubMed]
  508. Dolloff, N.G. Discovery platform for inhibitors of IgH gene enhancer activity. Cancer Biol. Ther. 2019, 20, 571–581. [Google Scholar] [CrossRef]
  509. Königsberg, R.; Ackermann, J.; Kaufmann, H.; Zojer, N.; Urbauer, E.; Krömer, E.; Jäger, U.; Gisslinger, H.; Schreiber, S.; Heinz, R.; et al. Deletions of chromosome 13q in monoclonal gammopathy of undetermined significance. Leukemia 2000, 14, 1975–1979. [Google Scholar] [CrossRef] [PubMed]
  510. Gertz, M.A.; Lacy, M.Q.; Dispenzieri, A.; Greipp, P.R.; Litzow, M.R.; Henderson, K.J.; Van Wier, S.A.; Ahmann, G.J.; Fonseca, R. Clinical implications of t(11;14)(q13;q32), t(4;14)(p16.3;q32), and -17p13 in myeloma patients treated with high-dose therapy. Blood 2005, 106, 2837–2840. [Google Scholar] [CrossRef] [PubMed]
  511. Fonseca, R.; Blood, E.; Rue, M.; Harrington, D.; Oken, M.M.; Kyle, R.A.; Dewald, G.W.; Van Ness, B.; Van Wier, S.A.; Henderson, K.J.; et al. Clinical and biologic implications of recurrent genomic aberrations in myeloma. Blood 2003, 101, 4569–4575. [Google Scholar] [CrossRef]
  512. Hanamura, I.; Stewart, J.P.; Huang, Y.; Zhan, F.; Santra, M.; Sawyer, J.R.; Hollmig, K.; Zangarri, M.; Pineda-Roman, M.; Van Rhee, F.; et al. Frequent gain of chromosome band 1q21 in plasma-cell dyscrasias detected by fluorescence in situ hybridization: Incidence increases from MGUS to relapsed myeloma and is related to prognosis and disease progression following tandem stem-cell transplantatio. Blood 2006, 108, 1724–1732. [Google Scholar] [CrossRef]
  513. Avet-Loiseau, H.; Attal, M.; Moreau, P.; Charbonnel, C.; Garban, F.; Hulin, C.; Leyvraz, S.; Michallet, M.; Yakoub-Agha, I.; Garderet, L.; et al. Genetic abnormalities and survival in multiple myeloma: The experience of the Intergroupe Francophone du Myélome. Blood 2007, 109, 3489–3495. [Google Scholar] [CrossRef] [PubMed]
  514. Avet-Loiseau, H.; Attal, M.; Campion, L.; Caillot, D.; Hulin, C.; Marit, G.; Stoppa, A.M.; Voillat, L.; Wetterwald, M.; Pegourie, B.; et al. Long-term analysis of the ifm 99 trials for myeloma: Cytogenetic abnormalities [t(4;14), del(17p), 1q gains] play a major role in defining long-term survival. J. Clin. Oncol. 2012, 30, 1949–1952. [Google Scholar] [CrossRef]
  515. Corre, J.; Perrot, A.; Caillot, D.; Belhadj, K.; Hulin, C.; Leleu, X.; Mohty, M.; Facon, T.; Buisson, L.; Do Souto, L.; et al. del(17p) without TP53 mutation confers a poor prognosis in intensively treated newly diagnosed patients with multiple myeloma. Blood 2021, 137, 1192–1195. [Google Scholar] [CrossRef] [PubMed]
  516. Chalopin, T.; Vallet, N.; Theisen, O.; Ochmann, M.; Tiab, M.; Godmer, P.; Barin, C.; Hérault, O.; Gyan, E.; Le Gouill, S.; et al. No survival improvement in patients with high-risk multiple myeloma harbouring del(17p) and/or t(4;14) over the two past decades. Br. J. Haematol. 2021, 194, 635–638. [Google Scholar] [CrossRef]
  517. Zhu, Y.X.; Shi, C.X.; Bruins, L.A.; Jedlowski, P.; Wang, X.; Kortüm, K.M.; Luo, M.; Ahmann, J.M.; Braggio, E.; Stewart, A.K. Loss of FAM46C promotes cell survival in myeloma. Cancer Res. 2017, 77, 4317–4327. [Google Scholar] [CrossRef] [PubMed]
  518. Kanasugi, J.; Hanamura, I.; Ota, A.; Karnan, S.; Lam, V.Q.; Mizuno, S.; Wahiduzzaman, M.; Rahman, M.L.; Hyodo, T.; Konishi, H.; et al. Biallelic loss of FAM46C triggers tumor growth with concomitant activation of Akt signaling in multiple myeloma cells. Cancer Sci. 2020, 111, 1663–1675. [Google Scholar] [CrossRef]
  519. Varma, A.; Sui, D.; Milton, D.R.; Tang, G.; Saini, N.; Hasan, O.; Mukherjee, A.; Joseph, J.J.; Bashir, Q.; Rondon, G.; et al. Outcome of Multiple Myeloma with Chromosome 1q Gain and 1p Deletion after Autologous Hematopoietic Stem Cell Transplantation: Propensity Score Matched Analysis. Biol. Blood Marrow Transplant. 2020, 26, 665–671. [Google Scholar] [CrossRef]
  520. Shah, G.L.; Landau, H.; Londono, D.; Devlin, S.M.; Kosuri, S.; Lesokhin, A.M.; Lendvai, N.; Hassoun, H.; Chung, D.J.; Koehne, G.; et al. Gain of chromosome 1q portends worse prognosis in multiple myeloma despite novel agent-based induction regimens and autologous transplantation. Leuk. Lymphoma 2017, 58, 1823–1831. [Google Scholar] [CrossRef]
  521. Chakraborty, R.; Gertz, M.A. + 1q: Amplifying the bad genes in myeloma. Leuk. 8 Lymphoma 2017, 58, 1771–1773. [Google Scholar] [CrossRef]
  522. Schmidt, T.M.; Barwick, B.G.; Joseph, N.; Heffner, L.T.; Hofmeister, C.C.; Bernal, L.; Dhodapkar, M.V.; Gupta, V.A.; Jaye, D.L.; Wu, J.; et al. Gain of Chromosome 1q is associated with early progression in multiple myeloma patients treated with lenalidomide, bortezomib, and dexamethasone. Blood Cancer J. 2019, 9, 94. [Google Scholar] [CrossRef]
  523. Sawyer, J.R.; Tricot, G.; Mattox, S.; Jagannath, S.; Barlogie, B. Jumping translocations of chromosome 1q in multiple myeloma: Evidence for a mechanism involving decondensation of pericentromeric heterochromatin. Blood 1998, 91, 1732–1741. [Google Scholar] [CrossRef]
  524. Jamet, D.; Marzin, Y.; Douet-Guilbert, N.; Morel, F.; Le Bris, M.-J.J.; Herry, A.; Banzakour, S.; Bourquard, P.; Morice, P.; Abgrall, J.F.; et al. Jumping translocations in multiple myeloma. Cancer Genet. Cytogenet. 2005, 161, 159–163. [Google Scholar] [CrossRef] [PubMed]
  525. Berger, R.; Bernard, O.A. Jumping translocations. Genes Chromosom. Cancer 2007, 46, 717–723. [Google Scholar] [CrossRef] [PubMed]
  526. Sawyer, J.R.; Tian, E.; Walker, B.A.; Wardell, C.; Lukacs, J.L.; Sammartino, G.; Bailey, C.; Schinke, C.D.; Thanendrarajan, S.; Davies, F.E.; et al. An acquired high-risk chromosome instability phenotype in multiple myeloma: Jumping 1q Syndrome. Blood Cancer J. 2019, 9, 62. [Google Scholar] [CrossRef]
  527. Philip, P.; Drivsholm, A.; Hansen, N.E.; Jensen, M.K.; Killmann, S.A. Chromosomes and survival in multiple myeloma. A banding study of 25 cases. Cancer Genet. Cytogenet. 1980, 2, 243–257. [Google Scholar] [CrossRef]
  528. Sawyer, J.R.; Tricot, G.; Lukacs, J.L.; Binz, R.L.; Tian, E.; Barlogie, B.; Shaughnessy, J. Genomic instability in multiple myeloma: Evidence for jumping segmental duplications of chromosome arm 1q. Genes Chromosom. Cancer 2005, 42, 95–106. [Google Scholar] [CrossRef]
  529. Aksenova, A.Y.; Mirkin, S.M. At the beginning of the end and in the middle of the beginning: Structure and maintenance of telomeric DNA repeats and interstitial telomeric sequences. Genes 2019, 10, 118. [Google Scholar] [CrossRef]
  530. Walker, B.A.; Mavrommatis, K.; Wardell, C.P.; Ashby, T.C.; Bauer, M.; Davies, F.; Rosenthal, A.; Wang, H.; Qu, P.; Hoering, A.; et al. A high-risk, Double-Hit, group of newly diagnosed myeloma identified by genomic analysis. Leukemia 2019, 33, 159–170. [Google Scholar] [CrossRef] [PubMed]
  531. Shaughnessy, J. Amplification and overexpression of CKS1B at chromosome band 1q21 is associated with reduced levels of p27 Kip1 and an aggressive clinical course in multiple myeloma. Hematology 2005, 10, 117–126. [Google Scholar] [CrossRef]
  532. Shi, W.; Huang, Q.; Xie, J.; Wang, H.; Yu, X.; Zhou, Y. CKS1B as Drug Resistance-Inducing Gene—A Potential Target to Improve Cancer Therapy. Front. Oncol. 2020, 10, 582451. [Google Scholar] [CrossRef]
  533. Hanamura, I. Gain/amplification of chromosome arm 1q21 in multiple myeloma. Cancers 2021, 13, 256. [Google Scholar] [CrossRef]
  534. An, G.; Xu, Y.; Shi, L.; Zhong, S.; Deng, S.; Xie, Z.; Sui, W.; Zhan, F.; Qiu, L. Chromosome 1q21 gains confer inferior outcomes in multiple myeloma treated with bortezomib but copy number variation and percentage of plasma cells involved have no additional prognostic value. Haematologica 2014, 99, 353–359. [Google Scholar] [CrossRef]
  535. Avet-Loiseau, H.; Li, J.Y.; Morineau, N.; Facon, T.; Brigaudeau, C.; Harousseau, J.L.; Grosbois, B.; Bataille, R. Monosomy 13 is associated with the transition of monoclonal gammopathy of undetermined significance to multiple myeloma. Intergroupe Francophone du Myélome. Blood 1999, 94, 2583–2589. [Google Scholar] [CrossRef]
  536. Shaughnessy, J.; Tian, E.; Sawyer, J.; Bumm, K.; Landes, R.; Badros, A.; Morris, C.; Tricot, G.; Epstein, J.; Barlogie, B. High incidence of chromosome 13 deletion in multiple myeloma detected by multiprobe interphase FISH. Blood 2000, 96, 1505–1511. [Google Scholar] [CrossRef] [PubMed]
  537. Chavan, S.S.; He, J.; Tytarenko, R.; Deshpande, S.; Patel, P.; Bailey, M.; Stein, C.K.; Stephens, O.; Weinhold, N.; Petty, N.; et al. Bi-allelic inactivation is more prevalent at relapse in multiple myeloma, identifying RB1 as an independent prognostic marker. Blood Cancer J. 2017, 7, e535. [Google Scholar] [CrossRef] [PubMed]
  538. Lodé, L.; Eveillard, M.; Trichet, V.; Soussi, T.; Wuillème, S.; Richebourg, S.; Magrangeas, F.; Ifrah, N.; Campion, L.; Traullé, C.; et al. Mutations in TP53 are exclusively associated with del(17p) in multiple myeloma. Haematologica 2010, 95, 1973–1976. [Google Scholar] [CrossRef]
  539. Chin, M.; Sive, J.I.; Allen, C.; Roddie, C.; Chavda, S.J.; Smith, D.; Blombery, P.; Jones, K.; Ryland, G.L.; Popat, R.; et al. Prevalence and timing of TP53 mutations in del(17p) myeloma and effect on survival. Blood Cancer J. 2017, 7, e610. [Google Scholar] [CrossRef]
  540. Sewastianik, T.; Prochorec-Sobieszek, M.; Chapuy, B.; Juszczyński, P. MYC deregulation in lymphoid tumors: Molecular mechanisms, clinical consequences and therapeutic implications. Biochim. Biophys. Acta-Rev. Cancer 2014, 1846, 457–467. [Google Scholar] [CrossRef] [PubMed]
  541. Grisanzio, C.; Freedman, M.L. Chromosome 8q24-associated cancers and MYC. Genes and Cancer 2010, 1, 555–559. [Google Scholar] [CrossRef]
  542. Huppi, K.; Pitt, J.J.; Wahlberg, B.M.; Caplen, N.J. The 8q24 gene desert: An oasis of non-coding transcriptional activity. Front. Genet. 2012, 3, 69. [Google Scholar] [CrossRef]
  543. Ghoussaini, M.; Song, H.; Koessler, T.; Al Olama, A.A.; Kote-Jarai, Z.; Driver, K.E.; Pooley, K.A.; Ramus, S.J.; Kjaer, S.K.; Hogdall, E.; et al. Multiple loci with different cancer specificities within the 8q24 gene desert. J. Natl. Cancer Inst. 2008, 100, 962–966. [Google Scholar] [CrossRef] [PubMed]
  544. Chesi, M.; Robbiani, D.F.; Sebag, M.; Chng, W.J.; Affer, M.; Tiedemann, R.; Valdez, R.; Palmer, S.E.; Haas, S.S.; Stewart, A.K.; et al. AID-Dependent Activation of a MYC Transgene Induces Multiple Myeloma in a Conditional Mouse Model of Post-Germinal Center Malignancies. Cancer Cell 2008, 13, 167–180. [Google Scholar] [CrossRef] [PubMed]
  545. Avet-Loiseau, H.; Gerson, F.; Magrangeas, F.; Minvielle, S.; Harousseau, J.L.; Bataille, R. Rearrangements of the c-myc oncogene are present in 15% of primary human multiple myeloma tumors. Blood 2001, 98, 3082–3086. [Google Scholar] [CrossRef] [PubMed]
  546. Misund, K.; Keane, N.; Stein, C.K.; Asmann, Y.W.; Day, G.; Welsh, S.; Van Wier, S.A.; Riggs, D.L.; Ahmann, G.; Chesi, M.; et al. MYC dysregulation in the progression of multiple myeloma. Leukemia 2020, 34, 322–326. [Google Scholar] [CrossRef]
  547. Affer, M.; Chesi, M.; Chen, W.D.; Keats, J.J.; Demchenko, Y.N.; Tamizhmani, K.; Garbitt, V.M.; Riggs, D.L.; Brents, L.A.; Roschke, A.V.; et al. Promiscuous MYC locus rearrangements hijack enhancers but mostly super-enhancers to dysregulate MYC expression in multiple myeloma. Leukemia 2014, 28, 1725–1735. [Google Scholar] [CrossRef]
  548. Barwick, B.G.; Neri, P.; Bahlis, N.J.; Nooka, A.K.; Dhodapkar, M.V.; Jaye, D.L.; Hofmeister, C.C.; Kaufman, J.L.; Gupta, V.A.; Auclair, D.; et al. Multiple myeloma immunoglobulin lambda translocations portend poor prognosis. Nat. Commun. 2019, 10, 1911. [Google Scholar] [CrossRef] [PubMed]
  549. Seliger, S.; Geirhos, V.; Haferlach, T.; Kern, W.; Walter, W.; Meggendorfer, M.; Baer, C.; Stengel, A.; Haferlach, C. Comprehensive Analysis of MYC Translocations in Multiple Myeloma By Whole Genome Sequencing and Whole Transcriptome Sequencing. Blood 2019, 134, 1774. [Google Scholar] [CrossRef]
  550. Walker, B.A.; Wardell, C.P.; Brioli, A.; Boyle, E.; Kaiser, M.F.; Begum, D.B.; Dahir, N.B.; Johnson, D.C.; Ross, F.M.; Davies, F.E.; et al. Translocations at 8q24 juxtapose MYC with genes that harbor superenhancers resulting in overexpression and poor prognosis in myeloma patients. Blood Cancer J. 2014, 4, e191. [Google Scholar] [CrossRef] [PubMed]
  551. Mikulasova, A.; Ashby, C.; Tytarenko, R.G.; Qu, P.; Rosenthal, A.; Dent, J.A.; Ryan, K.R.; Bauer, M.A.; Wardell, C.P.; Hoering, A.; et al. Microhomology-mediated end joining drives complex rearrangements and overexpression of MYC and PVT1 in multiple myeloma. Haematologica 2020, 105, 1055–1060. [Google Scholar] [CrossRef]
  552. Shaffer, A.L.; Emre, N.C.T.; Lamy, L.; Ngo, V.N.; Wright, G.; Xiao, W.; Powell, J.; Dave, S.; Yu, X.; Zhao, H.; et al. IRF4 addiction in multiple myeloma. Nature 2008, 454, 226–231. [Google Scholar] [CrossRef]
  553. Walker, B.A.; Wardell, C.P.; Johnson, D.C.; Kaiser, M.F.; Begum, D.B.; Dahir, N.B.; Ross, F.M.; Davies, F.E.; Gonzalez, D.; Morgan, G.J. Characterization of IGH locus breakpoints in multiple myeloma indicates a subset of translocations appear to occur in pregerminal center B cells. Blood 2013, 121, 3413–3419. [Google Scholar] [CrossRef]
  554. Jimenez, C.; Jara-Acevedo, M.; Corchete, L.A.L.A.; Castillo, D.; Ordonez, G.R.; Sarasquete, M.E.M.E.; Puig, N.; Martinez-Lopez, J.; Prieto-Conde, M.I.M.I.; Garcia-Alvarez, M.; et al. A Next-Generation Sequencing Strategy for Evaluating the Most Common Genetic Abnormalities in Multiple Myeloma. J. Mol. Diagnostics 2017, 19, 99–106. [Google Scholar] [CrossRef] [PubMed]
  555. Fear, D.J.; McCloskey, N.; O’Connor, B.; Felsenfeld, G.; Gould, H.J. Transcription of Ig Germline Genes in Single Human B Cells and the Role of Cytokines in Isotype Determination. J. Immunol. 2004, 173, 4529–4538. [Google Scholar] [CrossRef] [PubMed]
  556. Jumper, M.D.; Splawski, J.B.; Lipsky, P.E.; Meek, K. Ligation of CD40 induces sterile transcripts of multiple Ig H chain isotypes in human B cells. J. Immunol. 1994, 152, 438–445. [Google Scholar]
  557. Pène, J.; Gauchat, J.-F.; Lécart, S.; Drouet, E.; Guglielmi, P.; Boulay, V.; Delwail, A.; Foster, D.; Lecron, J.-C.; Yssel, H. Cutting Edge: IL-21 Is a Switch Factor for the Production of IgG 1 and IgG 3 by Human B Cells. J. Immunol. 2004, 172, 5154–5157. [Google Scholar] [CrossRef]
  558. Smith, C.I.E.; Baskin, B.; Pattersson, E.; Hammarstrom, L.; Islam, K.B. In Vivo Switching: Identification of Germline Transcripts For Human IgA. Adv. Mucosal Immunol. Part A 1995, 371A, 9–13. [Google Scholar]
  559. Fujieda, S.; Zhang, K.; Saxon, A. IL-4 plus CD40 monoclonal antibody induces human B cells gamma subclass-specific isotype switch: Switching to gamma 1, gamma 3, and gamma 4, but not gamma 2. J. Immunol. 1995, 155, 2318–2328. [Google Scholar] [PubMed]
  560. Splawski, J.B.; Fu, S.M.; Lipsky, P.E. Immunoregulatory role of CD40 in human B cell differentiation. J. Immunol. 1993, 150, 1276–1285. [Google Scholar] [PubMed]
  561. Boumendjel, A.; Tawk, L.; Malefijt, R. de W.; Boulay, V.; Yssel, H.; Pène, J. IL-27 induces the production of IgG1 by human B cells. Eur. Cytokine Netw. 2006, 17, 281–289. [Google Scholar]
  562. Lundgren, M.; Persson, U.; Larsson, P.; Magnusson, C.; Smith, C.I.E.; Hammarström, L.; Severinson, E. Interleukin 4 induces synthesis of IgE and IgG4 in human B cells. Eur. J. Immunol. 1989, 19, 1311–1315. [Google Scholar] [CrossRef]
  563. Cerutti, A.; Zan, H.; Schaffer, A.; Bergsagel, L.; Harindranath, N.; Max, E.E.; Casali, P. CD40 Ligand and Appropriate Cytokines Induce Switching to IgG, IgA, and IgE and Coordinated Germinal Center and Plasmacytoid Phenotypic Differentiation in a Human Monoclonal IgM+IgD+ B Cell Line. J. Immunol. 1998, 160, 2145–2157. [Google Scholar]
  564. Stavnezer-Nordgren, J.; Sirlin, S. Specificity of immunoglobulin heavy chain switch correlates with activity of germline heavy chain genes prior to switching. EMBO J. 1986, 5, 95–102. [Google Scholar] [CrossRef]
  565. Brière, F.; Servet-Delprat, C.; Bridon, J.M.; Saint-Remy, J.M.; Banchereau, J. Human interleukin 10 induces naive surface immunoglobulin D+ (sIgD+) B cells to secrete IgG1 and IgG3. J. Exp. Med. 1994, 179, 757–762. [Google Scholar] [CrossRef] [PubMed]
  566. Fujieda, S.; Saxon, A.; Zhang, K.E. Direct evidence that γ1 and γ3 switching in human B cells is interleukin-10 dependent. Mol. Immunol. 1996, 33, 1335–1343. [Google Scholar] [CrossRef]
  567. De Jong, B.G.; Ijspeert, H.; Marques, L.; Van Der Burg, M.; Van Dongen, J.J.M.; Loos, B.G.; Van Zelm, M.C. Human IgG2- and IgG4-expressing memory B cells display enhanced molecular and phenotypic signs of maturity and accumulate with age. Immunol. Cell Biol. 2017, 95, 744–752. [Google Scholar] [CrossRef] [PubMed]
  568. Horns, F.; Vollmers, C.; Croote, D.; Mackey, S.F.; Swan, G.E.; Dekker, C.L.; Davis, M.M.; Quake, S.R. Lineage tracing of human B cells reveals the in vivo landscape of human antibody class switching. Elife 2016, 5, e16578. [Google Scholar] [CrossRef]
  569. Jackson, K.J.L.; Wang, Y.; Collins, A.M. Human immunoglobulin classes and subclasses show variability in VDJ gene mutation levels. Immunol. Cell Biol. 2014, 92, 729–733. [Google Scholar] [CrossRef]
  570. Kitaura, K.; Yamashita, H.; Ayabe, H.; Shini, T.; Matsutani, T.; Suzuki, R. Different somatic hypermutation levels among antibody subclasses disclosed by a new next-generation sequencing-based antibody repertoire analysis. Front. Immunol. 2017, 8, 389. [Google Scholar] [CrossRef]
  571. Dayal, S.; Nedbal, J.; Hobson, P.; Cooper, A.M.; Gould, H.J.; Gellert, M.; Felsenfeld, G.; Fear, D.J. High resolution analysis of the chromatin landscape of the igE switch region in human B cells. PLoS ONE 2011, 6, e24571. [Google Scholar] [CrossRef]
  572. Nambu, Y.; Sugai, M.; Gonda, H.; Lee, C.G.; Katakai, T.; Agata, Y.; Yokota, Y.; Shimizu, A. Transcription-Coupled Events Associating with Immunoglobulin Switch Region Chromatin. Science 2003, 302, 2137–2140. [Google Scholar] [CrossRef]
  573. Li, Z.; Luo, Z.; Scharff, M.D. Differential regulation of histone acetylation and generation of mutations in switch regions is associated with Ig class switching. Proc. Natl. Acad. Sci. USA 2004, 101, 15428–15433. [Google Scholar] [CrossRef] [PubMed]
  574. Wang, L.; Whang, N.; Wuerffel, R.; Kenter, A.L. AID-dependent histone acetylation is detected in immunoglobulin S regions. J. Exp. Med. 2006, 203, 215–226. [Google Scholar] [CrossRef] [PubMed]
  575. Chowdhury, M.; Forouhi, O.; Dayal, S.; McCloskey, N.; Gould, H.J.; Felsenfeld, G.; Fear, D.J. Analysis of intergenic transcription and histone modification across the human immunoglobulin heavy-chain locus. Proc. Natl. Acad. Sci. USA 2008, 105, 15872–15877. [Google Scholar] [CrossRef]
  576. Wang, L.; Wuerffel, R.; Feldman, S.; Khamlichi, A.A.; Kenter, A.L. S region sequence, RNA polymerase II, and histone modifications create chromatin accessibility during class switch recombination. J. Exp. Med. 2009, 206, 1817–1830. [Google Scholar] [CrossRef] [PubMed]
  577. Kuang, F.L.; Luo, Z.; Scharff, M.D. Modifications Are Associated With Class Switch Recombination. Proc. Natl. Acad. Sci. 2009, 106, 5288–5293. [Google Scholar] [CrossRef]
  578. Perlot, T.; Alt, F.W.; Bassing, C.H.; Sun, H.; Pinaud, E. Elucidation of IgH intronic enhancer functions via germ-line deletion. Proc. Natl. Acad. Sci. USA 2005, 102, 14362–14367. [Google Scholar] [CrossRef] [PubMed]
  579. Vincent-Fabert, C.; Fiancette, R.; Pinaud, E.; Truffinet, V.; Cogné, N.; Cogné, M.; Denizot, Y. Genomic deletion of the whole IgH 3′ regulatory region (hs3a, hs1,2, hs3b, and hs4) dramatically affects class switch recombination and Ig secretion to all isotypes. Blood 2010, 116, 1895–1898. [Google Scholar] [CrossRef]
  580. Pinaud, E.; Marquet, M.; Fiancette, R.; Péron, S.; Vincent-Fabert, C.; Denizot, Y.; Cogné, M. The IgH locus 3’ regulatory region: Pulling the strings from behind. Adv. Immunol. 2011, 110, 27–70. [Google Scholar]
  581. Birshtein, B.K. Epigenetic regulation of individual modules of the immunoglobulin heavy chain locus 3’ regulatory region. Front. Immunol. 2014, 5, 163. [Google Scholar] [CrossRef]
  582. Sepulveda, M.A.; Emelyanov, A.V.; Birshtein, B.K. NF-κB and Oct-2 Synergize to Activate the Human 3′ Igh hs4 Enhancer in B Cells. J. Immunol. 2004, 172, 1054–1064. [Google Scholar] [CrossRef]
  583. Sepulveda, M.A.; Garrett, F.E.; Price-Whelan, A.; Birshtein, B.K. Comparative analysis of human and mouse 3′ Igh regulatory regions identifies distinctive structural features. Mol. Immunol. 2005, 42, 605–615. [Google Scholar] [CrossRef] [PubMed]
  584. Dalloul, I.; Boyer, F.; Dalloul, Z.; Pignarre, A.; Caron, G.; Fest, T.; Chatonnet, F.; Delaloy, C.; Durandy, A.; Jeannet, R.; et al. Locus suicide recombination actively occurs on the functionally rearranged IgH allele in B-cells from inflamed human lymphoid tissues. PLOS Genet. 2019, 15, e1007721. [Google Scholar] [CrossRef] [PubMed]
  585. Le Noir, S.; Bonaud, A.; Hervé, B.; Baylet, A.; Boyer, F.; Lecardeur, S.; Oruc, Z.; Sirac, C.; Cogné, M. IgH 3’ regulatory region increases ectopic class switch recombination. PLOS Genet. 2021, 17, e1009288. [Google Scholar] [CrossRef] [PubMed]
  586. Kuzin, I.I.; Bagaeva, L.; Young, F.M.; Bottaro, A. Requirement for Enhancer Specificity in Immunoglobulin Heavy Chain Locus Regulation. J. Immunol. 2008, 180, 7443–7450. [Google Scholar] [CrossRef]
  587. Kim, A.; Han, L.; Santiago, G.E.; Verdun, R.E.; Yu, K. Class-Switch Recombination in the Absence of the IgH 3′ Regulatory Region. J. Immunol. 2016, 197, 2930–2935. [Google Scholar] [CrossRef]
  588. Dauba, A.; Khamlichi, A.A. Long-Range Control of Class Switch Recombination by Transcriptional Regulatory Elements. Front. Immunol. 2021, 12, 3591. [Google Scholar] [CrossRef]
  589. Yu, K.; Lieber, M.R. Current insights into the mechanism of mammalian immunoglobulin class switch recombination. Crit. Rev. Biochem. Mol. Biol. 2019, 54, 333–351. [Google Scholar] [CrossRef]
  590. Stavnezer, J.; Schrader, C.E. IgH Chain Class Switch Recombination: Mechanism and Regulation. J. Immunol. 2014, 193, 5370–5378. [Google Scholar] [CrossRef]
  591. Montefiori, L.; Wuerffel, R.; Roqueiro, D.; Lajoie, B.; Guo, C.; Gerasimova, T.; De, S.; Wood, W.; Becker, K.G.; Dekker, J.; et al. Extremely Long-Range Chromatin Loops Link Topological Domains to Facilitate a Diverse Antibody Repertoire. Cell Rep. 2016, 14, 896–906. [Google Scholar] [CrossRef]
  592. Hill, L.; Ebert, A.; Jaritz, M.; Wutz, G.; Nagasaka, K.; Tagoh, H.; Kostanova-Poliakova, D.; Schindler, K.; Sun, Q.; Bönelt, P.; et al. Wapl repression by Pax5 promotes V gene recombination by Igh loop extrusion. Nature 2020, 584, 142–147. [Google Scholar] [CrossRef]
  593. Haarhuis, J.H.I.; van der Weide, R.H.; Blomen, V.A.; Yáñez-Cuna, J.O.; Amendola, M.; van Ruiten, M.S.; Krijger, P.H.L.; Teunissen, H.; Medema, R.H.; van Steensel, B.; et al. The Cohesin Release Factor WAPL Restricts Chromatin Loop Extension. Cell 2017, 169, 693–707.e14. [Google Scholar] [CrossRef]
  594. Wuerffel, R.; Wang, L.; Grigera, F.; Manis, J.; Selsing, E.; Perlot, T.; Alt, F.W.; Cogne, M.; Pinaud, E.; Kenter, A.L. S-S Synapsis during Class Switch Recombination Is Promoted by Distantly Located Transcriptional Elements and Activation-Induced Deaminase. Immunity 2007, 27, 711–722. [Google Scholar] [CrossRef] [PubMed]
  595. Oudinet, C.; Braikia, F.Z.; Dauba, A.; Khamlichi, A.A. Mechanism and regulation of class switch recombination by IgH transcriptional control elements. In Advances in Immunology; Academic Press Inc.: AEAmsterdam, The Netherlands, 2020; Volume 147. [Google Scholar]
  596. Chatterjee, S.; Ju, Z.; Hassan, R.; Volpi, S.A.; Emelyanov, A.V.; Birshtein, B.K. Dynamic changes in binding of immunoglobulin heavy chain 3′ regulatory region to protein factors during class switching. J. Biol. Chem. 2011, 286, 29303–29312. [Google Scholar] [CrossRef] [PubMed]
  597. Kenter, A.L.; Feldman, S.; Wuerffel, R.; Achour, I.; Wang, L.; Kumar, S. Three-dimensional architecture of the IgH locus facilitates class switch recombination. Ann. New York Acad. Sci. 2012, 1267, 86–94. [Google Scholar] [CrossRef]
  598. Degner, S.C.; Verma-Gaur, J.; Wong, T.P.; Bossen, C.; Iverson, G.M.; Torkamani, A.; Vettermann, C.; Lin, Y.C.; Ju, Z.; Schulz, D.; et al. CCCTC-binding factor (CTCF) and cohesin influence the genomic architecture of the Igh locus and antisense transcription in pro-B cells. Proc. Natl. Acad. Sci. USA 2011, 108, 9566–9571. [Google Scholar] [CrossRef]
  599. Zhang, X.; Yoon, H.S.; Chapdelaine-Williams, A.M.; Kyritsis, N.; Alt, F.W. Physiological role of the 3′IgH CBEs super-anchor in antibody class switching. Proc. Natl. Acad. Sci. USA 2021, 118, 2024392118. [Google Scholar] [CrossRef]
  600. Zhang, X.; Zhang, Y.; Ba, Z.; Kyritsis, N.; Casellas, R.; Alt, F.W. Fundamental roles of chromatin loop extrusion in antibody class switching. Nature 2019, 575, 385–389. [Google Scholar] [CrossRef] [PubMed]
  601. Ciccone, D.N.; Namiki, Y.; Chen, C.; Morshead, K.B.; Wood, A.L.; Johnston, C.M.; Morris, J.W.; Wang, Y.; Sadreyev, R.; Corcoran, A.E.; et al. The murine IgH locus contains a distinct DNA sequence motif for the chromatin regulatory factor CTCF. J. Biol. Chem. 2019, 294, 13580–13592. [Google Scholar] [CrossRef] [PubMed]
  602. Schwab, K.R.; Patel, S.R.; Dressler, G.R. Role of PTIP in Class Switch Recombination and Long-Range Chromatin Interactions at the Immunoglobulin Heavy Chain Locus. Mol. Cell. Biol. 2011, 31, 1503–1511. [Google Scholar] [CrossRef]
  603. Daniel, J.A.; Santos, M.A.; Wang, Z.; Zang, C.; Schwab, K.R.; Jankovic, M.; Filsuf, D.; Chen, H.T.; Gazumyan, A.; Yamane, A.; et al. PTIP promotes chromatin changes critical for immunoglobulin class switch recombination. Science 2010, 329, 917–923. [Google Scholar] [CrossRef]
  604. Giambra, V.; Volpi, S.; Emelyanov, A.V.; Pflugh, D.; Bothwell, A.L.M.; Norio, P.; Fan, Y.; Ju, Z.; Skoultchi, A.I.; Hardy, R.R.; et al. Pax5 and Linker Histone H1 Coordinate DNA Methylation and Histone Modifications in the 3′ Regulatory Region of the Immunoglobulin Heavy Chain Locus. Mol. Cell. Biol. 2008, 28, 6123–6133. [Google Scholar] [CrossRef]
  605. Minnich, M.; Tagoh, H.; Bönelt, P.; Axelsson, E.; Fischer, M.; Cebolla, B.; Tarakhovsky, A.; Nutt, S.L.; Jaritz, M.; Busslinger, M. Multifunctional role of the transcription factor Blimp-1 in coordinating plasma cell differentiation. Nat. Immunol. 2016, 17, 331–343. [Google Scholar] [CrossRef]
  606. Hauser, J.; Grundström, C.; Kumar, R.; Grundström, T. Regulated localization of an AID complex with E2A, PAX5 and IRF4 at the Igh locus. Mol. Immunol. 2016, 80, 78–90. [Google Scholar] [CrossRef]
  607. Grundström, C.; Kumar, A.; Priya, A.; Negi, N.; Grundström, T. ETS1 and PAX5 transcription factors recruit AID to Igh DNA. Eur. J. Immunol. 2018, 48, 1687–1697. [Google Scholar] [CrossRef]
  608. Perlot, T.; Li, G.; Alt, F.W. Antisense transcripts from immunoglobulin heavy-chain locus V(D)J and switch regions. Proc. Natl. Acad. Sci. USA 2008, 105, 3843–3848. [Google Scholar] [CrossRef] [PubMed]
  609. Osipovich, O.A.; Subrahmanyam, R.; Pierce, S.; Sen, R.; Oltz, E.M. Cutting Edge: SWI/SNF Mediates Antisense Igh Transcription and Locus-Wide Accessibility in B Cell Precursors. J. Immunol. 2009, 183, 1509–1513. [Google Scholar] [CrossRef] [PubMed]
  610. Yen, W.F.; Sharma, R.; Cols, M.; Lau, C.M.; Chaudhry, A.; Chowdhury, P.; Yewdell, W.T.; Vaidyanathan, B.; Sun, A.; Coffre, M.; et al. Distinct Requirements of CHD4 during B Cell Development and Antibody Response. Cell Rep. 2019, 27, 1472–1486.e5. [Google Scholar] [CrossRef]
  611. Lavery, W.J.; Barski, A.; Wiley, S.; Schorry, E.K.; Lindsley, A.W. KMT2C/D COMPASS complex-associated diseases [KCDCOM-ADs]: An emerging class of congenital regulopathies. Clin. Epigenetics 2020, 12, 10. [Google Scholar] [CrossRef] [PubMed]
  612. Eckner, R.; Yao, T.P.; Oldread, E.; Livingston, D.M. Interaction and functional collaboration of p300/CBP and bHLH proteins in muscle and B-cell differentiation. Genes Dev. 1996, 10, 2478–2490. [Google Scholar] [CrossRef] [PubMed]
  613. Pei, H.; Wu, X.; Liu, T.; Yu, K.; Jelinek, D.F.; Lou, Z. The Histone Methyltransferase MMSET Regulates Class Switch Recombination. J. Immunol. 2013, 190, 756–763. [Google Scholar] [CrossRef]
  614. Pefanis, E.; Wang, J.; Rothschild, G.; Lim, J.; Chao, J.; Rabadan, R.; Economides, A.N.; Basu, U. Noncoding RNA transcription targets AID to divergently transcribed loci in B cells. Nature 2014, 514, 389–393. [Google Scholar] [CrossRef]
  615. Medvedovic, J.; Ebert, A.; Tagoh, H.; Busslinger, M. Pax5: A Master Regulator of B Cell Development and Leukemogenesis. In Advances in Immunology; Academic Press Inc.: AEAmsterdam, The Netherlands, 2011; Volume 111, pp. 179–206. [Google Scholar]
  616. O’Brien, P.; Morin, P.; Ouellette, R.J.; Robichaud, G.A. The Pax-5 gene: A pluripotent regulator of B-cell differentiation and cancer disease. Cancer Res. 2011, 71, 7345–7350. [Google Scholar] [CrossRef]
  617. Okuyama, K.; Strid, T.; Kuruvilla, J.; Somasundaram, R.; Cristobal, S.; Smith, E.; Prasad, M.; Fioretos, T.; Lilljebjörn, H.; Soneji, S.; et al. PAX5 is part of a functional transcription factor network targeted in lymphoid leukemia. PLoS Genet. 2019, 15, e1008280. [Google Scholar] [CrossRef] [PubMed]
  618. Foltz, S.M.; Gao, Q.; Yoon, C.J.; Sun, H.; Yao, L.; Li, Y.; Jayasinghe, R.G.; Cao, S.; King, J.; Kohnen, D.R.; et al. Evolution and structure of clinically relevant gene fusions in multiple myeloma. Nat. Commun. 2020, 11, 2666. [Google Scholar] [CrossRef]
  619. Rustad, E.H.; Yellapantula, V.D.; Glodzik, D.; Maclachlan, K.H.; Diamond, B.; Boyle, E.M.; Ashby, C.; Blaney, P.; Gundem, G.; Hultcrantz, M.; et al. Revealing the Impact of Structural Variants in Multiple Myeloma. Blood Cancer Discov. 2020, 1, 258–273. [Google Scholar] [CrossRef] [PubMed]
  620. Qian, J.; Wang, Q.; Dose, M.; Pruett, N.; Kieffer-Kwon, K.R.; Resch, W.; Liang, G.; Tang, Z.; Mathé, E.; Benner, C.; et al. B cell super-enhancers and regulatory clusters recruit AID tumorigenic activity. Cell 2014, 159, 1524–1537. [Google Scholar] [CrossRef] [PubMed]
  621. Borson, N.D.; Lacy, M.Q.; Wettstein, P.J. Altered mRNA expression of Pax5 and Blimp-1 in B cells in multiple myeloma. Blood 2002, 100, 4629–4639. [Google Scholar] [CrossRef]
  622. Fuxa, M.; Skok, J.; Souabni, A.; Salvagiotto, G.; Roldan, E.; Busslinger, M. Pax5 induces V-to-DJ rearrangements and locus contraction of the immunoglobulin heavy-chain gene. Genes Dev. 2004, 18, 411–422. [Google Scholar] [CrossRef] [PubMed]
  623. Nutt, S.L.; Urbánek, P.; Relink, A.; Busslinger, M. Essential functions of Pax5 (BSAP) in pro-B cell development: Difference between fetal and adult B lymphopoiesis and reduced V-to-DJ recombination at the IgH locus. Genes Dev. 1997, 11, 476–491. [Google Scholar] [CrossRef]
  624. Waters, S.H.; Saikh, K.U.; Stavnezer, J. A B-cell-specific nuclear protein that binds to DNA sites 5’ to immunoglobulin S alpha tandem repeats is regulated during differentiation. Mol. Cell. Biol. 1989, 9, 5594–5601. [Google Scholar]
  625. Liao, F.; Giannini, S.L.; Birshtein, B.K. A nuclear DNA-binding protein expressed during early stages of B cell differentation interacts with diverse segments within and 3’ of the Ig H chain gene cluster. J. Immunol. 1992, 148, 2909–2917. [Google Scholar] [PubMed]
  626. Xu, L.; Kim, M.G.; Marcu, K.B. Properties of B cell stage specific and ubiquitous nuclear factors binding to immunoglobulin heavy chain gene switch regions. Int. Immunol. 1992, 4, 875–887. [Google Scholar] [CrossRef]
  627. Kosak, S.T.; Skok, J.A.; Medina, K.L.; Riblet, R.; Le Beau, M.M.; Fisher, A.G.; Singh, H. Subnuclear compartmentalization of immunoglobulin loci during lymphocyte development. Science 2002, 296, 158–162. [Google Scholar] [CrossRef]
  628. Zullo, J.M.; Demarco, I.A.; Piqué-Regi, R.; Gaffney, D.J.; Epstein, C.B.; Spooner, C.J.; Luperchio, T.R.; Bernstein, B.E.; Pritchard, J.K.; Reddy, K.L.; et al. DNA sequence-dependent compartmentalization and silencing of chromatin at the nuclear lamina. Cell 2012, 149, 1474–1487. [Google Scholar] [CrossRef] [PubMed]
  629. Bowen, A.J.; Corcoran, A.E. How chromatin remodelling allows shuffling of immunoglobulin heavy chain genes. Mol. Biosyst. 2008, 4, 790–798. [Google Scholar] [CrossRef] [PubMed]
  630. Corcoran, A.E. The epigenetic role of non-coding RNA transcription and nuclear organization in immunoglobulin repertoire generation. Semin. Immunol. 2010, 22, 353–361. [Google Scholar] [CrossRef]
  631. Pichugin, A.; Iarovaia, O.V.; Gavrilov, A.; Sklyar, I.; Barinova, N.; Barinov, A.; Ivashkin, E.; Caron, G.; Aoufouchi, S.; Razin, S.V.; et al. The IGH locus relocalizes to a “recombination compartment” in the perinucleolar region of differentiating B-lymphocytes. Oncotarget 2017, 8, 40079–40089. [Google Scholar] [CrossRef]
  632. Seifert, M.; Scholtysik, R.; Küppers, R. Origin and pathogenesis of B cell lymphomas. In Methods in Molecular Biology; Humana Press Inc.: Totowa, NJ, USA, 2019; Volume 1956, pp. 1–33. [Google Scholar]
  633. Ramiro, A.R.; Jankovic, M.; Eisenreich, T.; Difilippantonio, S.; Chen-Kiang, S.; Muramatsu, M.; Honjo, T.; Nussenzweig, A.; Nussenzweig, M.C. AID is required for c-myc/IgH chromosome translocations in vivo. Cell 2004, 118, 431–438. [Google Scholar] [CrossRef]
  634. Ghetti, M.; Vannini, I.; Storlazzi, C.T.; Martinelli, G.; Simonetti, G. Linear and circular PVT1 in hematological malignancies and immune response: Two faces of the same coin. Mol. Cancer 2020, 19, 69. [Google Scholar] [CrossRef] [PubMed]
  635. Derderian, C.; Orunmuyi, A.T.; Oluwabunmi Olapade-Olaopa, E.; Ogunwobi, O.O. PVT1 signaling is a mediator of cancer progression. Front. Oncol. 2019, 9, 502. [Google Scholar] [CrossRef]
  636. Cho, S.W.; Xu, J.; Sun, R.; Mumbach, M.R.; Carter, A.C.; Chen, Y.G.; Yost, K.E.; Kim, J.; He, J.; Nevins, S.A.; et al. Promoter of lncRNA Gene PVT1 Is a Tumor-Suppressor DNA Boundary Element. Cell 2018, 173, 1398–1412.e22. [Google Scholar] [CrossRef]
  637. Osborne, C.S.; Chakalova, L.; Mitchell, J.A.; Horton, A.; Wood, A.L.; Bolland, D.J.; Corcoran, A.E.; Fraser, P. Myc dynamically and preferentially relocates to a transcription factory occupied by Igh. PLoS Biol. 2007, 5, 1763–1772. [Google Scholar] [CrossRef]
  638. Strongin, D.E.; Groudine, M.; Politz, J.C.R. Nucleolar tethering mediates pairing between the IgH and Myc loci. Nucleus 2014, 5, 474–481. [Google Scholar] [CrossRef]
  639. Iarovaia, O.V.; Ioudinkova, E.S.; Razin, S.V.; Vassetzky, Y.S. Role of the Nucleolus in Rearrangements of the IGH Locus. Mol. Biol. (Mosk) 2018, 52, 210–219. [Google Scholar] [CrossRef]
  640. Sklyar, I.V.; Pichugin, A.M.; Razin, S.V.; Vassetzky, E.S.; Iarovaia, O.V. Nuclear localization of translocation partners in differentiating B-cells. Dokl. Biochem. Biophys. 2015, 464, 312–314. [Google Scholar] [CrossRef]
  641. Sklyar, I.; Iarovaia, O.V.; Gavrilov, A.A.; Pichugin, A.; Germini, D.; Tsfasman, T.; Caron, G.; Fest, T.; Lipinski, M.; Razin, S.V.; et al. Distinct Patterns of Colocalization of the CCND1 and CMYC Genes with Their Potential Translocation Partner IGH at Successive Stages of B-Cell Differentiation. J. Cell. Biochem. 2016, 117, 1506–1510. [Google Scholar] [CrossRef]
  642. Kaur, G.; Gupta, R.; Mathur, N.; Rani, L.; Kumar, L.; Sharma, A.; Singh, V.; Gupta, A.; Sharma, O.D. Clinical impact of chromothriptic complex chromosomal rearrangements in newly diagnosed multiple myeloma. Leuk. Res. 2019, 76, 58–64. [Google Scholar] [CrossRef]
  643. Smetana, J.; Oppelt, J.; Štork, M.; Pour, L.; Kuglík, P. Chromothripsis 18 in multiple myeloma patient with rapid extramedullary relapse. Mol. Cytogenet. 2018, 11, 7. [Google Scholar] [CrossRef] [PubMed]
  644. Magrangeas, F.; Avet-Loiseau, H.; Munshi, N.C.; Minvielle, S. Chromothripsis identifies a rare and aggressive entity among newly diagnosed multiple myeloma patients. Blood 2011, 118, 675–678. [Google Scholar] [CrossRef] [PubMed]
  645. Li, Y.; Roberts, N.D.; Wala, J.A.; Shapira, O.; Schumacher, S.E.; Kumar, K.; Khurana, E.; Waszak, S.; Korbel, J.O.; Haber, J.E.; et al. Patterns of somatic structural variation in human cancer genomes. Nature 2020, 578, 112–121. [Google Scholar] [CrossRef] [PubMed]
  646. Baca, S.C.; Prandi, D.; Lawrence, M.S.; Mosquera, J.M.; Romanel, A.; Drier, Y.; Park, K.; Kitabayashi, N.; MacDonald, T.Y.; Ghandi, M.; et al. Punctuated evolution of prostate cancer genomes. Cell 2013, 153, 666–677. [Google Scholar] [CrossRef] [PubMed]
  647. Marcozzi, A.; Pellestor, F.; Kloosterman, W.P. The genomic characteristics and origin of chromothripsis. In Methods in Molecular Biology; Humana Press Inc.: Totowa, NJ, USA, 2018; Volume 1769, pp. 3–19. [Google Scholar]
  648. Campbell, P.J.; Getz, G.; Korbel, J.O.; Stuart, J.M.; Jennings, J.L.; Stein, L.D.; Perry, M.D.; Nahal-Bose, H.K.; Ouellette, B.F.F.; Li, C.H.; et al. Pan-cancer analysis of whole genomes. Nature 2020, 578, 82–93. [Google Scholar]
  649. Kramara, J.; Osia, B.; Malkova, A. Break-Induced Replication: The Where, The Why, and The How. Trends Genet. 2018, 34, 518–531. [Google Scholar] [CrossRef]
  650. Kockler, Z.W.; Osia, B.; Lee, R.; Musmaker, K.; Malkova, A. Repair of DNA Breaks by Break-Induced Replication. Annu. Rev. Biochem. 2021, 90, 165–191. [Google Scholar] [CrossRef] [PubMed]
  651. Anand, R.P.; Tsaponina, O.; Greenwell, P.W.; Lee, C.S.; Du, W.; Petes, T.D.; Haber, J.E. Chromosome rearrangements via template switching between diverged repeated sequences. Genes Dev. 2014, 28, 2394–2406. [Google Scholar] [CrossRef]
  652. Rodgers, K.; McVey, M. Error-Prone Repair of DNA Double-Strand Breaks. J. Cell. Physiol. 2016, 231, 15–24. [Google Scholar] [CrossRef]
  653. Zhang, F.; Khajavi, M.; Connolly, A.M.; Towne, C.F.; Batish, S.D.; Lupski, J.R. The DNA replication FoSTeS/MMBIR mechanism can generate genomic, genic and exonic complex rearrangements in humans. Nat. Genet. 2009, 41, 849–853. [Google Scholar] [CrossRef] [PubMed]
  654. Holland, A.J.; Cleveland, D.W. Chromoanagenesis and cancer: Mechanisms and consequences of localized, complex chromosomal rearrangements. Nat. Med. 2012, 18, 1630–1638. [Google Scholar] [CrossRef]
  655. Hastings, P.J.; Ira, G.; Lupski, J.R. A Microhomology-Mediated Break-Induced Replication Model for the Origin of Human Copy Number Variation. PLoS Genet. 2009, 5, e1000327. [Google Scholar] [CrossRef]
  656. Stephens, P.J.; Greenman, C.D.; Fu, B.; Yang, F.; Bignell, G.R.; Mudie, L.J.; Pleasance, E.D.; Lau, K.W.; Beare, D.; Stebbings, L.A.; et al. Massive genomic rearrangement acquired in a single catastrophic event during cancer development. Cell 2011, 144, 27–40. [Google Scholar] [CrossRef]
  657. Lee, K.J.; Lee, K.H.; Yoon, K.-A.; Sohn, J.Y.; Lee, E.; Lee, H.; Eom, H.-S.; Kong, S.-Y. Chromothripsis in Treatment Resistance in Multiple Myeloma. Genom. Inform. 2017, 15, 87–97. [Google Scholar] [CrossRef]
  658. Maciejowski, J.; Li, Y.; Bosco, N.; Campbell, P.J.; De Lange, T. Chromothripsis and Kataegis Induced by Telomere Crisis. Cell 2015, 163, 1641–1654. [Google Scholar] [CrossRef]
  659. Mardin, B.R.; Drainas, A.P.; Waszak, S.M.; Weischenfeldt, J.; Isokane, M.; Stütz, A.M.; Raeder, B.; Efthymiopoulos, T.; Buccitelli, C.; Segura-Wang, M.; et al. A cell-based model system links chromothripsis with hyperploidy. Mol. Syst. Biol. 2015, 11, 828. [Google Scholar] [CrossRef]
  660. Li, Y.; Schwab, C.; Ryan, S.L.; Papaemmanuil, E.; Robinson, H.M.; Jacobs, P.; Moorman, A.V.; Dyer, S.; Borrow, J.; Griffiths, M.; et al. Constitutional and somatic rearrangement of chromosome 21 in acute lymphoblastic leukaemia. Nature 2014, 508, 98–102. [Google Scholar] [CrossRef]
  661. Garsed, D.W.; Marshall, O.J.; Corbin, V.D.A.; Hsu, A.; DiStefano, L.; Schröder, J.; Li, J.; Feng, Z.P.; Kim, B.W.; Kowarsky, M.; et al. The Architecture and Evolution of Cancer Neochromosomes. Cancer Cell 2014, 26, 653–667. [Google Scholar] [CrossRef]
  662. Liu, S.; Kwon, M.; Mannino, M.; Yang, N.; Renda, F.; Khodjakov, A.; Pellman, D. Nuclear envelope assembly defects link mitotic errors to chromothripsis. Nature 2018, 561, 551–555. [Google Scholar] [CrossRef] [PubMed]
  663. Zhang, C.Z.; Spektor, A.; Cornils, H.; Francis, J.M.; Jackson, E.K.; Liu, S.; Meyerson, M.; Pellman, D. Chromothripsis from DNA damage in micronuclei. Nature 2015, 522, 179–184. [Google Scholar] [CrossRef] [PubMed]
  664. Crasta, K.; Ganem, N.J.; Dagher, R.; Lantermann, A.B.; Ivanova, E.V.; Pan, Y.; Nezi, L.; Protopopov, A.; Chowdhury, D.; Pellman, D. DNA breaks and chromosome pulverization from errors in mitosis. Nature 2012, 482, 53–58. [Google Scholar] [CrossRef]
  665. Ly, P.; Brunner, S.F.; Shoshani, O.; Kim, D.H.; Lan, W.; Pyntikova, T.; Flanagan, A.M.; Behjati, S.; Page, D.C.; Campbell, P.J.; et al. Chromosome segregation errors generate a diverse spectrum of simple and complex genomic rearrangements. Nat. Genet. 2019, 51, 705–715. [Google Scholar] [CrossRef] [PubMed]
  666. Ye, C.J.; Sharpe, Z.; Alemara, S.; Mackenzie, S.; Liu, G.; Abdallah, B.; Horne, S.; Regan, S.; Heng, H.H. Micronuclei and genome chaos: Changing the system inheritance. Genes 2019, 10, 366. [Google Scholar] [CrossRef] [PubMed]
  667. Maciejowski, J.; De Lange, T. Telomeres in cancer: Tumour suppression and genome instability. Nat. Rev. Mol. Cell Biol. 2017, 18, 175–186. [Google Scholar] [CrossRef] [PubMed]
  668. Cleal, K.; Jones, R.E.; Grimstead, J.W.; Hendrickson, E.A.; Baird, D.M. Chromothripsis during telomere crisis is independent of NHEJ, and consistent with a replicative origin. Genome Res. 2019, 29, 737–749. [Google Scholar] [CrossRef]
  669. Cheng, C.; Zhou, Y.; Li, H.; Xiong, T.; Li, S.; Bi, Y.; Kong, P.; Wang, F.; Cui, H.; Li, Y.; et al. Whole-Genome Sequencing Reveals Diverse Models of Structural Variations in Esophageal Squamous Cell Carcinoma. Am. J. Hum. Genet. 2016, 98, 256–274. [Google Scholar] [CrossRef]
  670. Nik-Zainal, S.; Van Loo, P.; Wedge, D.C.; Alexandrov, L.B.; Greenman, C.D.; Lau, K.W.; Raine, K.; Jones, D.; Marshall, J.; Ramakrishna, M.; et al. The life history of 21 breast cancers. Cell 2012, 149, 994–1007. [Google Scholar] [CrossRef] [PubMed]
  671. Piazza, A.; Wright, W.D.; Heyer, W.D. Multi-invasions Are Recombination Byproducts that Induce Chromosomal Rearrangements. Cell 2017, 170, 760–773.e15. [Google Scholar] [CrossRef]
  672. Cottliar, A.; Pedrazzini, E.; Corrado, C.; Engelberger, M.I.; Narbaitz, M.; Slavutsky, I. Telomere shortening in patients with plasma cell disorders. Eur. J. Haematol. 2003, 71, 334–340. [Google Scholar] [CrossRef]
  673. de la Guardia, R.D.; Catalina, P.; Panero, J.; Elosua, C.; Pulgarin, A.; López, M.B.; Ayllón, V.; Ligero, G.; Slavutsky, I.; Leone, P.E. Expression profile of telomere-associated genes in multiple myeloma. J. Cell. Mol. Med. 2012, 16, 3009–3021. [Google Scholar] [CrossRef]
  674. Hyatt, S.; Jones, R.E.; Heppel, N.H.; Grimstead, J.W.; Fegan, C.; Jackson, G.H.; Hills, R.; Allan, J.M.; Pratt, G.; Pepper, C.; et al. Telomere length is a critical determinant for survival in multiple myeloma. Br. J. Haematol. 2017, 178, 94–98. [Google Scholar] [CrossRef]
  675. Wu, K.D.; Orme, L.M.; Shaughnessy, J.; Jacobson, J.; Barlogie, B.; Moore, M.A.S. Telomerase and telomere length in multiple myeloma: Correlations with disease heterogeneity, cytogenetic status, and overall survival. Blood 2003, 101, 4982–4989. [Google Scholar] [CrossRef]
  676. Panero, J.; Arbelbide, J.; Fantl, D.B.; Rivello, H.G.; Kohan, D.; Slavutsky, I. Altered mRNA expression of telomere-associated genes in monoclonal gammopathy of undetermined significance and multiple myeloma. Mol. Med. 2010, 16, 471–478. [Google Scholar] [CrossRef] [PubMed]
  677. Shiratsuchi, M.; Muta, K.; Abe, Y.; Motomura, S.; Taguchi, F.; Takatsuki, H.; Uike, N.; Umemura, T.; Nawata, H.; Nishimura, J. Clinical significance of telomerase activity in multiple myeloma. Cancer 2002, 94, 2232–2238. [Google Scholar] [CrossRef] [PubMed]
  678. Nergadze, S.G.; Santagostino, M.A.; Salzano, A.; Mondello, C.; Giulotto, E. Contribution of telomerase RNA retrotranscription to DNA double-strand break repair during mammalian genome evolution. Genome Biol. 2007, 8, R260. [Google Scholar] [CrossRef]
  679. Marzec, P.; Armenise, C.; Pérot, G.; Roumelioti, F.-M.M.; Basyuk, E.; Gagos, S.; Chibon, F.; Déjardin, J. Nuclear-Receptor-Mediated Telomere Insertion Leads to Genome Instability in ALT Cancers. Cell 2015, 160, 913–927. [Google Scholar] [CrossRef]
  680. Onozawa, M.; Zhang, Z.; Kim, Y.J.; Goldberg, L.; Varga, T.; Bergsagel, P.L.; Kuehl, W.M.; Aplan, P.D. Repair of DNA double-strand breaks by templated nucleotide sequence insertions derived from distant regions of the genome. Proc. Natl. Acad. Sci. USA 2014, 111, 7729–7734. [Google Scholar] [CrossRef] [PubMed]
  681. Azzalin, C.M.; Mucciolo, E.; Bertoni, L.; Giulotto, E. Fluorescence in situ hybridization with a synthetic (T2AG3)n polynucleotide detects several intrachromosomal telomere-like repeats on human chromosomes. Cytogenet. Cell Genet. 1997, 78, 112–115. [Google Scholar] [CrossRef]
  682. Aksenova, A.Y.; Greenwell, P.W.; Dominska, M.; Shishkin, A.A.; Kim, J.C.; Petes, T.D.; Mirkin, S.M. Genome rearrangements caused by interstitial telomeric sequences in yeast. Proc. Natl. Acad. Sci. USA 2013, 110, 19866–19871. [Google Scholar] [CrossRef] [PubMed]
  683. Kilburn, A.E.; Shea, M.J.; Sargent, R.G.; Wilson, J.H. Insertion of a telomere repeat sequence into a mammalian gene causes chromosome instability. Mol. Cell. Biol. 2001, 21, 126–135. [Google Scholar] [CrossRef]
  684. Bosco, N.; de Lange, T. A TRF1-controlled common fragile site containing interstitial telomeric sequences. Chromosoma 2012, 121, 465–474. [Google Scholar] [CrossRef]
  685. Wood, A.M.; Danielsen, J.M.R.; Lucas, C.A.; Rice, E.L.; Scalzo, D.; Shimi, T.; Goldman, R.D.; Smith, E.D.; Le Beau, M.M.; Kosak, S.T. TRF2 and lamin A/C interact to facilitate the functional organization of chromosome ends. Nat. Commun. 2014, 5, 5467. [Google Scholar] [CrossRef]
  686. Wood, A.M.; Laster, K.; Rice, E.L.; Kosak, S.T. A beginning of the end: New insights into the functional organization of telomeres. Nucleus 2015, 6, 172–178. [Google Scholar] [CrossRef]
  687. Klewes, L.; Vallente, R.; Dupas, E.; Brand, C.; Grün, D.; Guffei, A.; Sathitruangsak, C.; Awe, J.A.; Kuzyk, A.; Lichtensztejn, D.; et al. Three-dimensional nuclear telomere organization in multiple myeloma. Transl. Oncol. 2013, 6, 749–756. [Google Scholar] [CrossRef] [PubMed]
  688. Yu, P.L.I.; Wang, R.R.; Johnston, G.; Wang, Y.; Tammur, P.; Tamm, A.; Punab, M.; Rangel-Pozzo, A.; Mai, S. Distinct Nuclear Organization of Telomeresand Centromeres in Monoclonal Gammopathyof Undetermined Significance and Multiple Myeloma. Cells 2019, 8, 723. [Google Scholar] [CrossRef] [PubMed]
  689. Campa, D.; Martino, A.; Varkonyi, J.; Lesueur, F.; Jamroziak, K.; Landi, S.; Jurczyszyn, A.; Marques, H.; Andersen, V.; Jurado, M.; et al. Risk of multiple myeloma is associated with polymorphisms within telomerase genes and telomere length. Int. J. Cancer 2015, 136, E351–E358. [Google Scholar] [CrossRef] [PubMed]
  690. Hosnijeh, F.S.; Matullo, G.; Russo, A.; Guarrera, S.; Modica, F.; Nieters, A.; Overvad, K.; Guldberg, P.; Tjønneland, A.; Canzian, F.; et al. Prediagnostic telomere length and risk of B-cell lymphoma-Results from the EPIC cohort study. Int. J. Cancer 2014, 135, 2910–2917. [Google Scholar] [CrossRef]
  691. Went, M.; Cornish, A.J.; Law, P.J.; Kinnersley, B.; van Duin, M.; Weinhold, N.; Försti, A.; Hansson, M.; Sonneveld, P.; Goldschmidt, H.; et al. Search for multiple myeloma risk factors using Mendelian randomization. Blood Adv. 2020, 4, 2172–2179. [Google Scholar] [CrossRef]
  692. Jones, A.M.; Beggs, A.D.; Carvajal-Carmona, L.; Farrington, S.; Tenesa, A.; Walker, M.; Howarth, K.; Ballereau, S.; Hodgson, S.V.; Zauber, A.; et al. TERC polymorphisms are associated both with susceptibility to colorectal cancer and with longer telomeres. Gut 2012, 61, 248–254. [Google Scholar] [CrossRef] [PubMed]
  693. Houlston, R.S.; Cheadle, J.; Dobbins, S.E.; Tenesa, A.; Jones, A.M.; Howarth, K.; Spain, S.L.; Broderick, P.; Domingo, E.; Farrington, S.; et al. Meta-analysis of three genome-wide association studies identifies susceptibility loci for colorectal cancer at 1q41, 3q26.2, 12q13.13 and 20q13.33. Nat. Genet. 2010, 42, 973–979. [Google Scholar] [CrossRef] [PubMed]
  694. Codd, V.; Nelson, C.P.; Albrecht, E.; Mangino, M.; Deelen, J.; Buxton, J.L.; Hottenga, J.J.; Fischer, K.; Esko, T.; Surakka, I.; et al. Identification of seven loci affecting mean telomere length and their association with disease. Nat. Genet. 2013, 45, 422–427. [Google Scholar] [CrossRef]
  695. Speedy, H.E.; Di Bernardo, M.C.; Sava, G.P.; Dyer, M.J.S.; Holroyd, A.; Wang, Y.; Sunter, N.J.; Mansouri, L.; Juliusson, G.; Smedby, K.E.; et al. A genome-wide association study identifies multiple susceptibility loci for chronic lymphocytic leukemia. Nat. Genet. 2014, 46, 56–60. [Google Scholar] [CrossRef]
  696. Michalek, J.E.; Kepa, A.; Vincent, J.; Frissa, S.; Goodwin, L.; Hotopf, M.; Hatch, S.L.; Breen, G.; Powell, T.R. Genetic predisposition to advanced biological ageing increases risk for childhood-onset recurrent major depressive disorder in a large UK sample. J. Affect. Disord. 2017, 213, 207–213. [Google Scholar] [CrossRef]
  697. Bojesen, S.E.; Pooley, K.A.; Johnatty, S.E.; Beesley, J.; Michailidou, K.; Tyrer, J.P.; Edwards, S.L.; Pickett, H.A.; Shen, H.C.; Smart, C.E.; et al. Multiple independent variants at the TERT locus are associated with telomere length and risks of breast and ovarian cancer. Nat. Genet. 2013, 45, 371–384. [Google Scholar] [CrossRef] [PubMed]
  698. Coutts, F.; Palmos, A.B.; Duarte, R.R.R.; de Jong, S.; Lewis, C.M.; Dima, D.; Powell, T.R. The polygenic nature of telomere length and the anti-ageing properties of lithium. Neuropsychopharmacology 2019, 44, 757–765. [Google Scholar] [CrossRef]
  699. Palm, W.; de Lange, T. How shelterin protects mammalian telomeres. Annu. Rev. Genet. 2008, 42, 301–334. [Google Scholar] [CrossRef] [PubMed]
  700. Speedy, H.E.; Kinnersley, B.; Chubb, D.; Broderick, P.; Law, P.J.; Litchfield, K.; Jayne, S.; Dyer, M.J.S.; Dearden, C.; Follows, G.A.; et al. Germ line mutations in shelterin complex genes are associated with familial chronic lymphocytic leukemia. Blood 2016, 128, 2319–2326. [Google Scholar] [CrossRef] [PubMed]
  701. Panero, J.; Stanganelli, C.; Arbelbide, J.; Fantl, D.B.; Kohan, D.; García Rivello, H.; Rabinovich, G.A.; Slavutsky, I. Expression profile of shelterin components in plasma cell disorders. Clinical significance of POT1 overexpression. Blood Cells Mol. Dis. 2014, 52, 134–139. [Google Scholar] [CrossRef]
  702. Popuri, V.; Hsu, J.; Khadka, P.; Horvath, K.; Liu, Y.; Croteau, D.L.; Bohr, V.A. Human RECQL1 participates in telomere maintenance. Nucleic Acids Res. 2014, 42, 5671–5688. [Google Scholar] [CrossRef]
  703. Viziteu, E.; Klein, B.; Basbous, J.; Lin, Y.L.; Hirtz, C.; Gourzones, C.; Tiers, L.; Bruyer, A.; Vincent, L.; Grandmougin, C.; et al. RECQ1 helicase is involved in replication stress survival and drug resistance in multiple myeloma. Leukemia 2017, 31, 2104–2113. [Google Scholar] [CrossRef]
  704. Viziteu, E.; Kassambara, A.; Pasero, P.; Klein, B.; Moreaux, J. RECQ helicases are deregulated in hematological malignancies in association with a prognostic value. Biomark. Res. 2016, 4, 3. [Google Scholar] [CrossRef]
  705. Porro, A.; Feuerhahn, S.; Lingner, J. TERRA-reinforced association of LSD1 with MRE11 promotes processing of uncapped telomeres. Cell Rep. 2014, 6, 765–776. [Google Scholar] [CrossRef]
  706. Hirschi, A.; Martin, W.J.; Luka, Z.; Loukachevitch, L.V.; Reiter, N.J. G-quadruplex RNA binding and recognition by the lysine-specific histone demethylase-1 enzyme. RNA 2016, 22, 1250–1260. [Google Scholar] [CrossRef]
  707. Rahmanto, Y.S.; Jung, J.G.; Wu, R.C.; Kobayashi, Y.; Heaphy, C.M.; Meeker, A.K.; Wang, T.L.; Shih, I.M. Inactivating ARID1A tumor suppressor enhances TERT transcription and maintains telomere length in cancer cells. J. Biol. Chem. 2016, 291, 9690–9699. [Google Scholar] [CrossRef]
  708. Zhao, B.; Lin, J.; Rong, L.; Wu, S.; Deng, Z.; Fatkhutdinov, N.; Zundell, J.; Fukumoto, T.; Liu, Q.; Kossenkov, A.; et al. ARID1A promotes genomic stability through protecting telomere cohesion. Nat. Commun. 2019, 10, 4067. [Google Scholar] [CrossRef] [PubMed]
  709. Zhu, X.D.; Niedernhofer, L.; Kuster, B.; Mann, M.; Hoeijmakers, J.H.J.; De Lange, T. ERCC1/XPF Removes the 3′ Overhang from Uncapped Telomeres and Represses Formation of Telomeric DNA-Containing Double Minute Chromosomes. Mol. Cell 2003, 12, 1489–1498. [Google Scholar] [CrossRef]
  710. Tseng, C.K.; Wang, H.F.; Burns, A.M.M.; Schroeder, M.R.R.; Gaspari, M.; Baumann, P. Human Telomerase RNA Processing and Quality Control. Cell Rep. 2015, 13, 2232–2243. [Google Scholar] [CrossRef] [PubMed]
  711. Robinson, S.R.; Oliver, A.W.; Chevassut, T.J.; Newbury, S.F. The 3’ to 5’ exoribonuclease DIS3: From structure and mechanisms to biological functions and role in human disease. Biomolecules 2015, 5, 1515–1539. [Google Scholar] [CrossRef] [PubMed]
  712. Her, Y.R.; Chung, I.K. P300-mediated acetylation of TRF2 is required for maintaining functional telomeres. Nucleic Acids Res. 2013, 41, 2267–2283. [Google Scholar] [CrossRef]
  713. Cubiles, M.D.; Barroso, S.; Vaquero-Sedas, M.I.; Enguix, A.; Aguilera, A.; Vega-Palas, M.A. Epigenetic features of human telomeres. Nucleic Acids Res. 2018, 46, 2347–2355. [Google Scholar] [CrossRef]
  714. Shalem-Cohavi, N.; Beery, E.; Nordenberg, J.; Rozovski, U.; Raanani, P.; Lahav, M.; Uziel, O. The effects of proteasome inhibitors on telomerase activity and regulation in multiple myeloma cells. Int. J. Mol. Sci. 2019, 20, 2509. [Google Scholar] [CrossRef]
  715. Weiss, C.; Uziel, O.; Wolach, O.; Nordenberg, J.; Beery, E.; Bulvick, S.; Kanfer, G.; Cohen, O.; Ram, R.; Bakhanashvili, M.; et al. Differential downregulation of telomerase activity by bortezomib in multiple myeloma cells-multiple regulatory pathways in vitro and ex vivo. Br. J. Cancer 2012, 107, 1844–1852. [Google Scholar] [CrossRef]
  716. Frontzek, F.; Staiger, A.M.; Zapukhlyak, M.; Xu, W.; Bonzheim, I.; Borgmann, V.; Sander, P.; Baptista, M.J.; Heming, J.N.; Berning, P.; et al. Molecular and functional profiling identifies therapeutically targetable vulnerabilities in plasmablastic lymphoma. Nat. Commun. 2021, 12, 1–14. [Google Scholar] [CrossRef]
  717. Liu, Z.; Filip, I.; Gomez, K.; Engelbrecht, D.; Meer, S.; Lalloo, P.N.; Patel, P.; Perner, Y.; Zhao, J.; Wang, J.; et al. Genomic Characterization of HIV-Associated Plasmablastic Lymphoma Identifies Pervasive Mutations in the JAK–STAT Pathway. Blood Cancer Discov. 2020, 1, 112–125. [Google Scholar] [CrossRef]
  718. Leeman-Neill, R.J.; Soderquist, C.R.; Montanari, F.; Raciti, P.; Park, D.; Radeski, D.; Mansukhani, M.M.; Murty, V.V.; Hsiao, S.; Alobeid, B.; et al. Phenogenomic heterogeneity of post-transplant plasmablastic lymphomas. Haematologica 2020. online ahead of print. [Google Scholar]
  719. Paradzik, T.; Bandini, C.; Mereu, E.; Labrador, M.; Taiana, E.; Amodio, N.; Neri, A.; Piva, R. The landscape of signaling pathways and proteasome inhibitors combinations in multiple myeloma. Cancers 2021, 13, 1235. [Google Scholar] [CrossRef]
  720. Yang, Y.; Li, Y.; Gu, H.; Dong, M.; Cai, Z. Emerging agents and regimens for multiple myeloma. J. Hematol. Oncol. 2020, 13, 150. [Google Scholar] [CrossRef] [PubMed]
  721. Tamura, H.; Ishibashi, M.; Sunakawa, M.; Inokuchi, K. Immunotherapy for Multiple Myeloma. Cancers 2019, 11, 2009. [Google Scholar] [CrossRef]
  722. Tian, Z.; D’Arcy, P.; Wang, X.; Ray, A.; Tai, Y.T.; Hu, Y.; Carrasco, R.D.; Richardson, P.; Linder, S.; Chauhan, D.; et al. A novel small molecule inhibitor of deubiquitylating enzyme USP14 and UCHL5 induces apoptosis in multiple myeloma and overcomes bortezomib resistance. Blood 2014, 123, 706–716. [Google Scholar] [CrossRef]
  723. John, L.; Krauth, M.T.; Podar, K.; Raab, M.S. Pathway-directed therapy in multiple myeloma. Cancers 2021, 13, 1668. [Google Scholar] [CrossRef]
  724. Alagpulinsa, D.A.; Szalat, R.E.; Poznansky, M.C.; Shmookler Reis, R.J. Genomic Instability in Multiple Myeloma. Trends in Cancer 2020, 6, 858–873. [Google Scholar] [CrossRef]
  725. Taiana, E.; Cantafio, M.E.G.; Favasuli, V.K.; Bandini, C.; Viglietto, G.; Piva, R.; Neri, A.; Amodio, N. Genomic instability in multiple myeloma: A “non-coding rna” perspective. Cancers 2021, 13, 2127. [Google Scholar] [CrossRef]
  726. Caprio, C.; Sacco, A.; Giustini, V.; Roccaro, A.M. Epigenetic aberrations in multiple myeloma. Cancers 2020, 12, 2996. [Google Scholar] [CrossRef] [PubMed]
  727. White, B.S.; Lanc, I.; O’Neal, J.; Gupta, H.; Fulton, R.S.; Schmidt, H.; Fronick, C.; Belter, E.A., Jr.; Fiala, M.; King, J.; et al. A multiple myeloma-specific capture sequencing platform discovers novel translocations and frequent, risk-associated point mutations in IGLL5. Blood Cancer J. 2018, 8, 35. [Google Scholar] [CrossRef] [PubMed]
  728. Yellapantula, V.; Hultcrantz, M.; Rustad, E.H.; Wasserman, E.; Londono, D.; Cimera, R.; Ciardiello, A.; Landau, H.; Akhlaghi, T.; Mailankody, S.; et al. Comprehensive detection of recurring genomic abnormalities: A targeted sequencing approach for multiple myeloma. Blood Cancer J. 2019, 9, 101. [Google Scholar] [CrossRef] [PubMed]
  729. Buedts, L.; Smits, S.; Ameye, G.; Lehnert, S.; Ding, J.; Delforge, M.; Vermeesch, J.; Boeckx, N.; Tousseyn, T.; Michaux, L.; et al. Ultra-low depth sequencing of plasma cell DNA for the detection of copy number aberrations in multiple myeloma. Genes Chromosom. Cancer 2020, 59, 465–471. [Google Scholar] [CrossRef] [PubMed]
  730. Laganà, A.; Beno, I.; Melnekoff, D.; Leshchenko, V.; Madduri, D.; Ramdas, D.; Sanchez, L.; Niglio, S.; Perumal, D.; Kidd, B.A.; et al. Precision Medicine for Relapsed Multiple Myeloma on the Basis of an Integrative Multiomics Approach. JCO Precis. Oncol. 2018, 2018, PO.18.00019. [Google Scholar] [CrossRef]
  731. Jang, J.S.; Li, Y.; Mitra, A.K.; Bi, L.; Abyzov, A.; van Wijnen, A.J.; Baughn, L.B.; Van Ness, B.; Rajkumar, V.; Kumar, S.; et al. Molecular signatures of multiple myeloma progression through single cell RNA-Seq. Blood Cancer J. 2019, 9, 2. [Google Scholar] [CrossRef]
  732. Höllein, A.; Twardziok, S.; Walter, W.; Hutter, S.; Baer, C.; Hernandez-Sanchez, J.M.; Meggendorfer, M.; Haferlach, T.; Kern, W.; Haferlach, C.; et al. The combination of WGS and RNA-Seq is superior to conventional diagnostic tests in multiple myeloma: Ready for prime time? Cancer Genet. 2020, 242, 15–24. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Stages of MM development and their characteristics. MM is almost always preceded by a precancerous condition termed monoclonal gammopathy of undetermined significance, MGUS [5,6,7]. Smoldering multiple myeloma (SMM) is an intermediate stage between MGUS and MM [12,13]. MGUS is diagnosed when serum monoclonal protein (M-protein) is detected, but levels are lower than in MM; a 2–3-fold elevation of the number of clonal bone marrow plasma cells is usually found in MGUS compared to healthy individuals. In MGUS, end-organ damage, CRAB, attributed to the plasma cell proliferative disorder, is absent [1]. SMM is diagnosed when serum or urinary monoclonal protein rises and/or the number of clonal bone marrow plasma cells increases, but there is no MDE or amyloidosis [1]. There are germline risk alleles in several genes associated with familiar cases of MM and MGUS; also, numerous SNPs found in GWAS are associated with MM development risk. Telomere status is another factor that can affect MM development. Primary events can be found as early as MGUS and are represented by structural genome changes, gain of chromosomes, and mutations. The primary structural genomic changes are nearly equally represented by the two major groups: the IGH translocations and trisomies of several odd chromosomes (referred to as hyperdiploidy). Few other structural aberrations found as early as the MGUS stage are listed. While the disease progresses, more genomic changes accumulate. The events that are characteristic of SMM and MM, but not MGUS, are classified as secondary. They are represented by various structural genome changes including MYC translocations, accumulation of complex genome rearrangements (partially, since some of them can be seen at MGUS stage), mutations in various pathways, and AID/APOBEC-induced mutation burden. The endpoint of the scheme illustrates treatment-induced events. Some of them can modulate DNA repair and impact genome stability and mutation accumulation. Abbreviations: IMiDs—Immunomodulatory drugs; CRAB—an acronym for Calcium (elevated), Renal failure, Anemia, and Bone lesions, the most common symptoms of MM; MDE—myeloma-defining events (see text for more details); MRI—magnetic resonance imaging; BM—bone marrow; AID—activation-induced cytosine deaminase essential for class switch recombination (CSR) and somatic hypermutation (SHM); APOBEC—a family of cytosine deaminases that play an essential role in mutagenesis in cancer cells; GWAS—Genome-Wide Association Studies.
Figure 1. Stages of MM development and their characteristics. MM is almost always preceded by a precancerous condition termed monoclonal gammopathy of undetermined significance, MGUS [5,6,7]. Smoldering multiple myeloma (SMM) is an intermediate stage between MGUS and MM [12,13]. MGUS is diagnosed when serum monoclonal protein (M-protein) is detected, but levels are lower than in MM; a 2–3-fold elevation of the number of clonal bone marrow plasma cells is usually found in MGUS compared to healthy individuals. In MGUS, end-organ damage, CRAB, attributed to the plasma cell proliferative disorder, is absent [1]. SMM is diagnosed when serum or urinary monoclonal protein rises and/or the number of clonal bone marrow plasma cells increases, but there is no MDE or amyloidosis [1]. There are germline risk alleles in several genes associated with familiar cases of MM and MGUS; also, numerous SNPs found in GWAS are associated with MM development risk. Telomere status is another factor that can affect MM development. Primary events can be found as early as MGUS and are represented by structural genome changes, gain of chromosomes, and mutations. The primary structural genomic changes are nearly equally represented by the two major groups: the IGH translocations and trisomies of several odd chromosomes (referred to as hyperdiploidy). Few other structural aberrations found as early as the MGUS stage are listed. While the disease progresses, more genomic changes accumulate. The events that are characteristic of SMM and MM, but not MGUS, are classified as secondary. They are represented by various structural genome changes including MYC translocations, accumulation of complex genome rearrangements (partially, since some of them can be seen at MGUS stage), mutations in various pathways, and AID/APOBEC-induced mutation burden. The endpoint of the scheme illustrates treatment-induced events. Some of them can modulate DNA repair and impact genome stability and mutation accumulation. Abbreviations: IMiDs—Immunomodulatory drugs; CRAB—an acronym for Calcium (elevated), Renal failure, Anemia, and Bone lesions, the most common symptoms of MM; MDE—myeloma-defining events (see text for more details); MRI—magnetic resonance imaging; BM—bone marrow; AID—activation-induced cytosine deaminase essential for class switch recombination (CSR) and somatic hypermutation (SHM); APOBEC—a family of cytosine deaminases that play an essential role in mutagenesis in cancer cells; GWAS—Genome-Wide Association Studies.
Cancers 13 05949 g001
Figure 2. MM treatment outlines. The schemes align with the latest recommendations and studies and include available options for MM treatment as well as standards of care [87,88,90,91,103]. Action of some drugs on DNA integrity or pathways affecting DNA metabolism/repair is shown on the scheme. All patients can be subdivided into three categories based on frailty: “go-go”, “slow-go”, and “no-go” [111]. “Slow-go” and “no-go” categories are usually treated using the reduced intensity regimens or dose-adjusted regimens and are not shown on the scheme. Treatment for the “no-go” category excludes chemotherapeutic drugs and includes agents with low toxicity, mostly for palliative care [111]. The “go-go” category and to some extent the “slow-go” category are shown on the scheme. VRd * can be used as induction (initial) therapy for newly diagnosed and transplant-eligible patients with MM and is associated with a progression-free survival longer than 4 years and overall survival at 4 years higher than 80% [87,103]. The alternative scheme is VTd, which is slightly less efficient [103]. The addition of the monoclonal antibody daratumumab targeting CD38 to the standard VRd or VTd regimen (Quadruplet regimen, Dara-VRd *, or Dara-VTd) is recommended for high-risk patients [87]. DARA-Rd is also an important alternative to VRd in newly diagnosed MM [112]. Besides that, VCd can be used as an induction scheme [88]. High-dose melphalan and ASCT give a good response and provide the longest remission [87,102]. Single-agent lenalidomide, administered continuously until disease progression is considered the standard of care for maintenance, and bortezomib can be used for high-risk patients. In addition, ixazomib can be considered in place of bortezomib [87,103,113]. If the disease relapses, another ASCT can be tried for eligible patients [100,101,102,114,115]; new treatment regimens are usually considered, including quadruplet schemes and new drugs [87]. The alkylating agent cyclophosphamide can be used instead of immunomodulatory agents as treatment for refractory patients and in patients with renal dysfunction in combination with bortezomib and dexamethasone; this setting can also be used for induction therapy in ASCT-eligible patients [87,116,117,118,119,120,121,122,123,124,125]. Melphalan and prednisone-based therapy$ can be used for ASCT-ineligible patients and often is represented by VMP or DARA-VMP schemes [88]. Continuous therapy or initial therapy with maintenance can be used in a non-transplant setting [113]. A multi-regimen, including bortezomib, an immunomodulatory drug, dexamethasone, cytotoxic cisplatin, doxorubicin, cyclophosphamide, and etoposide, may be used for plasma cell leukemia or extramedullary disease [87]. Schemes that can be used for refractory to lenalidomide and/or bortezomib patients are also given in the center of the figure. Abbreviations: V—bortezomib, DARA—daratumumab, R—lenalidomide, T—thalidomide, POM —pomalidomide, d—dexamethasone; K—carfilzomib, I—ixazomib, E—elotuzumab, Isa—isatuximab, C—cyclophosphamide, P—prednisone; S—selinexor, Ven—venetoclax, ICL—intermolecular crosslinks in the DNA, FA/BRCA—pathway responsible for the repair of intermolecular crosslinks in the DNA, HR—homologous recombination, HRR—homologous recombination repair, DDR—DNA damage response, NER—nucleotide excision repair. *—Lenalidomide-containing regimens (e.g., VRd and Dara-VRd) are not yet approved by European Medicines Agency (EMA, EU) as induction for ASCT-eligible patients, although they offer good risk–benefit profiles and are widely used in the USA [87,88]. **—Bortezomib and ixazomib have not yet been approved by the EMA for maintenance after ASCT [88,113]. ***—Awaiting EMA approval. $—melphalan-containing regimens in this setting are not recommended in the USA due to concerns about SPMs and stem cells damage.
Figure 2. MM treatment outlines. The schemes align with the latest recommendations and studies and include available options for MM treatment as well as standards of care [87,88,90,91,103]. Action of some drugs on DNA integrity or pathways affecting DNA metabolism/repair is shown on the scheme. All patients can be subdivided into three categories based on frailty: “go-go”, “slow-go”, and “no-go” [111]. “Slow-go” and “no-go” categories are usually treated using the reduced intensity regimens or dose-adjusted regimens and are not shown on the scheme. Treatment for the “no-go” category excludes chemotherapeutic drugs and includes agents with low toxicity, mostly for palliative care [111]. The “go-go” category and to some extent the “slow-go” category are shown on the scheme. VRd * can be used as induction (initial) therapy for newly diagnosed and transplant-eligible patients with MM and is associated with a progression-free survival longer than 4 years and overall survival at 4 years higher than 80% [87,103]. The alternative scheme is VTd, which is slightly less efficient [103]. The addition of the monoclonal antibody daratumumab targeting CD38 to the standard VRd or VTd regimen (Quadruplet regimen, Dara-VRd *, or Dara-VTd) is recommended for high-risk patients [87]. DARA-Rd is also an important alternative to VRd in newly diagnosed MM [112]. Besides that, VCd can be used as an induction scheme [88]. High-dose melphalan and ASCT give a good response and provide the longest remission [87,102]. Single-agent lenalidomide, administered continuously until disease progression is considered the standard of care for maintenance, and bortezomib can be used for high-risk patients. In addition, ixazomib can be considered in place of bortezomib [87,103,113]. If the disease relapses, another ASCT can be tried for eligible patients [100,101,102,114,115]; new treatment regimens are usually considered, including quadruplet schemes and new drugs [87]. The alkylating agent cyclophosphamide can be used instead of immunomodulatory agents as treatment for refractory patients and in patients with renal dysfunction in combination with bortezomib and dexamethasone; this setting can also be used for induction therapy in ASCT-eligible patients [87,116,117,118,119,120,121,122,123,124,125]. Melphalan and prednisone-based therapy$ can be used for ASCT-ineligible patients and often is represented by VMP or DARA-VMP schemes [88]. Continuous therapy or initial therapy with maintenance can be used in a non-transplant setting [113]. A multi-regimen, including bortezomib, an immunomodulatory drug, dexamethasone, cytotoxic cisplatin, doxorubicin, cyclophosphamide, and etoposide, may be used for plasma cell leukemia or extramedullary disease [87]. Schemes that can be used for refractory to lenalidomide and/or bortezomib patients are also given in the center of the figure. Abbreviations: V—bortezomib, DARA—daratumumab, R—lenalidomide, T—thalidomide, POM —pomalidomide, d—dexamethasone; K—carfilzomib, I—ixazomib, E—elotuzumab, Isa—isatuximab, C—cyclophosphamide, P—prednisone; S—selinexor, Ven—venetoclax, ICL—intermolecular crosslinks in the DNA, FA/BRCA—pathway responsible for the repair of intermolecular crosslinks in the DNA, HR—homologous recombination, HRR—homologous recombination repair, DDR—DNA damage response, NER—nucleotide excision repair. *—Lenalidomide-containing regimens (e.g., VRd and Dara-VRd) are not yet approved by European Medicines Agency (EMA, EU) as induction for ASCT-eligible patients, although they offer good risk–benefit profiles and are widely used in the USA [87,88]. **—Bortezomib and ixazomib have not yet been approved by the EMA for maintenance after ASCT [88,113]. ***—Awaiting EMA approval. $—melphalan-containing regimens in this setting are not recommended in the USA due to concerns about SPMs and stem cells damage.
Cancers 13 05949 g002
Figure 3. A simplified scheme of B-cell development based on studies in humans and mice. The scheme illustrates the principal stages of B-cell development (predominantly of the B2 lineage) based on studies in mice and humans. Commitment to the B-cell lineage requires the action of several transcription factors (some of them are illustrated in the figure). PAX5, BACH2, PU.1, and IRF8 are expressed throughout B-cell development and are silenced in antibody-secreting cells [212,213,214]. Surrogate light chains, SLC (VpreB, and λ14.1), and Igα production are the hallmarks of the pro-B-cell stage. V(D)J rearrangement results in the synthesis of the immunoglobulin heavy chain (µH) with Cµ constant domain chain. The µH together with SLC and Igαβ form a pre-BCR receptor, which is required to pass the first essential quality control during B-cell development, after which a rearrangement at the IGL or IGK loci encoding the λ or κ light chains is initiated. IRF4, along with IRF8, interacts with PU.1 and binds to Ig κ 3′ enhancer and λ enhancers to modulate rearrangement of Ig light chains at the pre-B cell stage [213,215]. This permits the production of the conventional IgM molecules on the cell surface, which corresponds to the “Immature B-cell” stage. The IgM BCR passes rounds of controls to eliminate self-reactive molecules. Dual expression of the IgM BCR and IgD BCR on the cell surface identifies the “Mature B-cell” stage. The mature naïve B-cells migrate to the periphery into the secondary lymphoid organs, where they are subjected to antigen presentation and selection. Upon antigen stimulation, they might form special structures called Germinal Centers (GC), where the antigen-binding sites of the antibodies may be further adaptively altered in the course of SHM, allowing the production of high-affinity antibodies. Starting from this stage, the B-cell may become either a memory cell or the antibody-secreting plasma cell [212,216]. Another type of antibody-secreting cell is plasmablast, plasmablasts are cycling cells that are produced early in the immune response. IRF4, along with BLIMP1(PRDM1) and XBP1, are expressed in antibody-secreting plasma cells and play a role in plasma cell commitment. IRF4, BLIMP1, and XBP1 are often found mutated or dysregulated in MM (see Section 3 and Table S2). IRF4 is induced by BCR signaling and upregulates BLIMP1, which has positive feedback on IRF4 [212]. BLIMP1 is a master-regulator that represses PAX5 and BCL6 programs, induces the expression of XBP1, and turns the program toward plasma cell development [217]. IRF4 upregulates AICDA, encoding AID, and is essential for CSR. CSR is ceased through the BLIMP1-mediated pathway upon B-cell differentiation into plasmablasts or plasma cells [214,217].
Figure 3. A simplified scheme of B-cell development based on studies in humans and mice. The scheme illustrates the principal stages of B-cell development (predominantly of the B2 lineage) based on studies in mice and humans. Commitment to the B-cell lineage requires the action of several transcription factors (some of them are illustrated in the figure). PAX5, BACH2, PU.1, and IRF8 are expressed throughout B-cell development and are silenced in antibody-secreting cells [212,213,214]. Surrogate light chains, SLC (VpreB, and λ14.1), and Igα production are the hallmarks of the pro-B-cell stage. V(D)J rearrangement results in the synthesis of the immunoglobulin heavy chain (µH) with Cµ constant domain chain. The µH together with SLC and Igαβ form a pre-BCR receptor, which is required to pass the first essential quality control during B-cell development, after which a rearrangement at the IGL or IGK loci encoding the λ or κ light chains is initiated. IRF4, along with IRF8, interacts with PU.1 and binds to Ig κ 3′ enhancer and λ enhancers to modulate rearrangement of Ig light chains at the pre-B cell stage [213,215]. This permits the production of the conventional IgM molecules on the cell surface, which corresponds to the “Immature B-cell” stage. The IgM BCR passes rounds of controls to eliminate self-reactive molecules. Dual expression of the IgM BCR and IgD BCR on the cell surface identifies the “Mature B-cell” stage. The mature naïve B-cells migrate to the periphery into the secondary lymphoid organs, where they are subjected to antigen presentation and selection. Upon antigen stimulation, they might form special structures called Germinal Centers (GC), where the antigen-binding sites of the antibodies may be further adaptively altered in the course of SHM, allowing the production of high-affinity antibodies. Starting from this stage, the B-cell may become either a memory cell or the antibody-secreting plasma cell [212,216]. Another type of antibody-secreting cell is plasmablast, plasmablasts are cycling cells that are produced early in the immune response. IRF4, along with BLIMP1(PRDM1) and XBP1, are expressed in antibody-secreting plasma cells and play a role in plasma cell commitment. IRF4, BLIMP1, and XBP1 are often found mutated or dysregulated in MM (see Section 3 and Table S2). IRF4 is induced by BCR signaling and upregulates BLIMP1, which has positive feedback on IRF4 [212]. BLIMP1 is a master-regulator that represses PAX5 and BCL6 programs, induces the expression of XBP1, and turns the program toward plasma cell development [217]. IRF4 upregulates AICDA, encoding AID, and is essential for CSR. CSR is ceased through the BLIMP1-mediated pathway upon B-cell differentiation into plasmablasts or plasma cells [214,217].
Cancers 13 05949 g003
Figure 4. Localization of regions with specific mutational signatures in the multiple myeloma genomes. Mutational signatures identified in MM genomes are shown on panel (A). Panels (BD) schematically show locations of these signatures on the chromosomes. There is substantial variability in the types of translocations and the strength of each individual mutational signature in the individual patient genomes and within tumor subclones. Thus, the scheme represents “averaged” features of the mutational signatures. Deamination in CpG motifs (yellow boxes with green border) is scattered randomly in the genomes. Translocations involving the IGH locus can possess all four deaminase-related mutational signatures found in myeloma genomes. An example of translocation involving the IGH and MYC loci is shown (B). Both AID- and APOBEC3-related mutations can be found around the breakpoint. As with other translocations (panel C), APOBEC3-related mutations (green and blue) are happening due to the deamination of ssDNA resulting from the resection of double-stranded breaks. The amount of resection, the lifetime of resected ends, and availability of APOBEC3 proteins in the nucleus at the time of the translocation event should affect the length and density of the hypermutated regions. Clusters can span over Mb distances around the breakpoint, with most dense regions having inter-mutational distance less than 1 Kb. AID-related mutations (red and orange) are found not only in the IGH locus but also scattered throughout the genome, often found in the 5’-non-coding regions and generally in the beginnings of the genes (BD). This is a result of the off-target effects of SHM. These clusters are generally smaller than APOBEC3-related clusters. Genes may possess one of the AID-related signatures or both. Gene bodies are more likely to possess APOBEC3-related mutations (D). This effect occurs because, even though deamination of 5’-regions of the genes is a natural feature of all APOBEC proteins, there are no known trans-acting factors capable of recruiting APOBEC3 enzymes to the beginning of the genes or switch regions of the IGH locus, as happens in cases of AID in SHM and CSR. This scattered signature of APOBEC3 in the genome is different from classical kataegis in that there are no significant and dense mutational clusters and is more reminiscent of “omikli”, which has been recently described in other cancers [455].
Figure 4. Localization of regions with specific mutational signatures in the multiple myeloma genomes. Mutational signatures identified in MM genomes are shown on panel (A). Panels (BD) schematically show locations of these signatures on the chromosomes. There is substantial variability in the types of translocations and the strength of each individual mutational signature in the individual patient genomes and within tumor subclones. Thus, the scheme represents “averaged” features of the mutational signatures. Deamination in CpG motifs (yellow boxes with green border) is scattered randomly in the genomes. Translocations involving the IGH locus can possess all four deaminase-related mutational signatures found in myeloma genomes. An example of translocation involving the IGH and MYC loci is shown (B). Both AID- and APOBEC3-related mutations can be found around the breakpoint. As with other translocations (panel C), APOBEC3-related mutations (green and blue) are happening due to the deamination of ssDNA resulting from the resection of double-stranded breaks. The amount of resection, the lifetime of resected ends, and availability of APOBEC3 proteins in the nucleus at the time of the translocation event should affect the length and density of the hypermutated regions. Clusters can span over Mb distances around the breakpoint, with most dense regions having inter-mutational distance less than 1 Kb. AID-related mutations (red and orange) are found not only in the IGH locus but also scattered throughout the genome, often found in the 5’-non-coding regions and generally in the beginnings of the genes (BD). This is a result of the off-target effects of SHM. These clusters are generally smaller than APOBEC3-related clusters. Genes may possess one of the AID-related signatures or both. Gene bodies are more likely to possess APOBEC3-related mutations (D). This effect occurs because, even though deamination of 5’-regions of the genes is a natural feature of all APOBEC proteins, there are no known trans-acting factors capable of recruiting APOBEC3 enzymes to the beginning of the genes or switch regions of the IGH locus, as happens in cases of AID in SHM and CSR. This scattered signature of APOBEC3 in the genome is different from classical kataegis in that there are no significant and dense mutational clusters and is more reminiscent of “omikli”, which has been recently described in other cancers [455].
Cancers 13 05949 g004
Figure 5. The schematic map of the human IGH locus illustrating events occurring during B-cell development. V(D)J rearrangement and CSR are thought to be the two main sources of genomic instability leading to the IGH translocations in MM. Translocation breakpoints mapped in MM are shown as black circles under the corresponding elements of the IGH locus in the germline configuration on the top. The breakpoints’ coordinates were taken from [553,554]. Then they were located on the chromosome using the GRCh38.p13 reference assembly and were plotted on the map to illustrate their distribution relative to structural elements of the IGH locus. Most breakpoints concentrate around Eμ and Sμ as well as in other switch regions such as Sγ1, Sγ2, and Sγ3. Some breakpoints are scattered within V, D, and J regions. CSR depends on the transcription of the switch regions located upstream of most genes encoding the immunoglobulin constant regions. Germline transcription starts from TATA-less promoters located upstream of the small I-exons (not shown) and the switch regions. Transcription goes through the entire S-region and the corresponding CH gene. Isotype switching depends on the corresponding specific transcription (red and blue arrows) and is stimulated by external stimuli such as cytokines (IL-4, IL-21, IL-10, IL-27), CD40 ligands, or T cell-independent signals (e.g., lipopolysaccharide, LPS [555,556,557,558,559,560,561,562,563,564,565,566]). Several studies support the idea of sequential switching of immunoglobulin genes [567,568,569,570]. Primary and secondary switching routes are illustrated as thick, black arrows and dotted, thin arrows, correspondingly. It is generally believed that switch region chromatin structure plays a key role in promoting transcription-coupled AID attack. Changes in histone modifications were observed in switch regions upon cytokine stimulation [571,572,573,574,575,576,577].
Figure 5. The schematic map of the human IGH locus illustrating events occurring during B-cell development. V(D)J rearrangement and CSR are thought to be the two main sources of genomic instability leading to the IGH translocations in MM. Translocation breakpoints mapped in MM are shown as black circles under the corresponding elements of the IGH locus in the germline configuration on the top. The breakpoints’ coordinates were taken from [553,554]. Then they were located on the chromosome using the GRCh38.p13 reference assembly and were plotted on the map to illustrate their distribution relative to structural elements of the IGH locus. Most breakpoints concentrate around Eμ and Sμ as well as in other switch regions such as Sγ1, Sγ2, and Sγ3. Some breakpoints are scattered within V, D, and J regions. CSR depends on the transcription of the switch regions located upstream of most genes encoding the immunoglobulin constant regions. Germline transcription starts from TATA-less promoters located upstream of the small I-exons (not shown) and the switch regions. Transcription goes through the entire S-region and the corresponding CH gene. Isotype switching depends on the corresponding specific transcription (red and blue arrows) and is stimulated by external stimuli such as cytokines (IL-4, IL-21, IL-10, IL-27), CD40 ligands, or T cell-independent signals (e.g., lipopolysaccharide, LPS [555,556,557,558,559,560,561,562,563,564,565,566]). Several studies support the idea of sequential switching of immunoglobulin genes [567,568,569,570]. Primary and secondary switching routes are illustrated as thick, black arrows and dotted, thin arrows, correspondingly. It is generally believed that switch region chromatin structure plays a key role in promoting transcription-coupled AID attack. Changes in histone modifications were observed in switch regions upon cytokine stimulation [571,572,573,574,575,576,577].
Cancers 13 05949 g005
Figure 6. Factors affecting genomic stability and CSR at the IGH locus are implicated in MM development. A hypothetical scheme (based on studies in murine and human models) that includes factors essential for the IGH locus reorganization and CSR. Many of these factors are encoded by genes mutations in which predispose to MM development and genes that are recurrently mutated in MM. Development of MM may be linked to genomic instability at the IGH locus: IGH translocations are observed as early as MGUS, are frequent, and predispose to disease development since, in most cases, they fuse various oncogenes to powerful enhancers at the IGH locus. Since IGH translocations may be linked to aberrant CSR, this is evidence that MM initiation can be simply a result of aberrant CSR. CSR depends on the function of several genetic elements at the IGH locus, such as 3’RR, Eμ, and S-regions. It also depends on a number of factors that aid chromatin changes and transcription through S-regions, RNA maturation and processing, and AID recruitment. Transcription factors PAX5 and BCL6 play an essential role in the IGH locus contraction and remodeling. PAX5 is a master regulator that binds to the DNA, and recruits PTIP and MLL3/MLL4 methyltransferase complex to specific regions at the IGH locus, thus promoting chromatin changes and transcription initiation at S-regions [602,603]. PAX5, together with Linker histone H1 (green spheres with tails), influences DNA methylation and histone modifications at the IGH locus [604]. BCL6 binds the promoter region of γ1 GLT and may act by repressing the transcription at S-regions [316]. BLIMP1 and IRF4 bind to 3′RR and regulate transcription at IGH locus; BLIMP1 also represses transcription of AICDA and BCL6, whereas IRF4 activates AICDA expression [200,217,605]. AID co-localizes with several transcription factors, including PAX5 and IRF4, at the IGH locus, which might play a role in AID targeting to specific locations [606,607]. Chromatin-remodeling factors ARID1A and CHD4 facilitate nucleosome reorganization required for transcription through S-regions. ARID1A is a component of the SWI/SNF complex known to be involved in antisense transcription at the IGH locus [608,609]. CHD4 is a component of the NuRD complex, which binds to H3K9me3, an epigenetic mark present at the IGH locus during CSR. CHD4 is required for CSR and coimmunoprecipitates with AID in B-cells, and this interaction was proposed to aid AID to target S-regions [610]. LSD1 is a lysine-specific demethylase that associates with the NuRD complex and also interacts with BCL6 [315]. p300 and CBP are closely related acetyltransferases that play a role in transcription activation and enhancer regulation [611,612]. KDM6A acts in concert with MLL3/MLL4 and p300/CBP in the mediation of transcription initiation. MMSET is lysine methyltransferase, which is required for GLT transcription at S-regions [613]. In addition, MMSET promotes AID-mediated DNA breaks at the donor switch region [502]. DIS3 is a component of the exosome, the complex that facilitates 3′-5′ RNA processing and aids in AID recruitment to both DNA strands [331,332,614]. In addition to these mechanisms, the telomeric location of the IGH locus suggests that any events that affect telomeric length and structure or telomeric chromatin may influence the CSR and provoke chromosomal aberrations.
Figure 6. Factors affecting genomic stability and CSR at the IGH locus are implicated in MM development. A hypothetical scheme (based on studies in murine and human models) that includes factors essential for the IGH locus reorganization and CSR. Many of these factors are encoded by genes mutations in which predispose to MM development and genes that are recurrently mutated in MM. Development of MM may be linked to genomic instability at the IGH locus: IGH translocations are observed as early as MGUS, are frequent, and predispose to disease development since, in most cases, they fuse various oncogenes to powerful enhancers at the IGH locus. Since IGH translocations may be linked to aberrant CSR, this is evidence that MM initiation can be simply a result of aberrant CSR. CSR depends on the function of several genetic elements at the IGH locus, such as 3’RR, Eμ, and S-regions. It also depends on a number of factors that aid chromatin changes and transcription through S-regions, RNA maturation and processing, and AID recruitment. Transcription factors PAX5 and BCL6 play an essential role in the IGH locus contraction and remodeling. PAX5 is a master regulator that binds to the DNA, and recruits PTIP and MLL3/MLL4 methyltransferase complex to specific regions at the IGH locus, thus promoting chromatin changes and transcription initiation at S-regions [602,603]. PAX5, together with Linker histone H1 (green spheres with tails), influences DNA methylation and histone modifications at the IGH locus [604]. BCL6 binds the promoter region of γ1 GLT and may act by repressing the transcription at S-regions [316]. BLIMP1 and IRF4 bind to 3′RR and regulate transcription at IGH locus; BLIMP1 also represses transcription of AICDA and BCL6, whereas IRF4 activates AICDA expression [200,217,605]. AID co-localizes with several transcription factors, including PAX5 and IRF4, at the IGH locus, which might play a role in AID targeting to specific locations [606,607]. Chromatin-remodeling factors ARID1A and CHD4 facilitate nucleosome reorganization required for transcription through S-regions. ARID1A is a component of the SWI/SNF complex known to be involved in antisense transcription at the IGH locus [608,609]. CHD4 is a component of the NuRD complex, which binds to H3K9me3, an epigenetic mark present at the IGH locus during CSR. CHD4 is required for CSR and coimmunoprecipitates with AID in B-cells, and this interaction was proposed to aid AID to target S-regions [610]. LSD1 is a lysine-specific demethylase that associates with the NuRD complex and also interacts with BCL6 [315]. p300 and CBP are closely related acetyltransferases that play a role in transcription activation and enhancer regulation [611,612]. KDM6A acts in concert with MLL3/MLL4 and p300/CBP in the mediation of transcription initiation. MMSET is lysine methyltransferase, which is required for GLT transcription at S-regions [613]. In addition, MMSET promotes AID-mediated DNA breaks at the donor switch region [502]. DIS3 is a component of the exosome, the complex that facilitates 3′-5′ RNA processing and aids in AID recruitment to both DNA strands [331,332,614]. In addition to these mechanisms, the telomeric location of the IGH locus suggests that any events that affect telomeric length and structure or telomeric chromatin may influence the CSR and provoke chromosomal aberrations.
Cancers 13 05949 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Aksenova, A.Y.; Zhuk, A.S.; Lada, A.G.; Zotova, I.V.; Stepchenkova, E.I.; Kostroma, I.I.; Gritsaev, S.V.; Pavlov, Y.I. Genome Instability in Multiple Myeloma: Facts and Factors. Cancers 2021, 13, 5949. https://doi.org/10.3390/cancers13235949

AMA Style

Aksenova AY, Zhuk AS, Lada AG, Zotova IV, Stepchenkova EI, Kostroma II, Gritsaev SV, Pavlov YI. Genome Instability in Multiple Myeloma: Facts and Factors. Cancers. 2021; 13(23):5949. https://doi.org/10.3390/cancers13235949

Chicago/Turabian Style

Aksenova, Anna Y., Anna S. Zhuk, Artem G. Lada, Irina V. Zotova, Elena I. Stepchenkova, Ivan I. Kostroma, Sergey V. Gritsaev, and Youri I. Pavlov. 2021. "Genome Instability in Multiple Myeloma: Facts and Factors" Cancers 13, no. 23: 5949. https://doi.org/10.3390/cancers13235949

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