Previous Article in Journal
Polyphenols and Chronic Myeloid Leukemia: Emerging Therapeutic Opportunities
Previous Article in Special Issue
Diagnosis and Treatment of Langerhans Cell Sarcoma: A Case Report and Review of the Literature
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Multiple Myeloma and Precursor Plasma Cell Disorders: From Emerging Driver Mutations to Current and Future Therapeutic Strategies

by
Henry Sutanto
1,2,*,
Pradana Zaky Romadhon
3,4,*,
Vembi Rizky Fatmawati
1,2,
Alief Waitupu
1,2,
Bagus Aditya Ansharullah
1,2,
Betty Rachma
1,2,
Elisa Elisa
1,2,
Laras Pratiwi
1,2 and
Galih Januar Adytia
1,2
1
Internal Medicine Study Program, Department of Internal Medicine, Faculty of Medicine, Universitas Airlangga, Surabaya 60132, Indonesia
2
Department of Internal Medicine, Dr. Soetomo General Academic Hospital, Surabaya 60286, Indonesia
3
Division of Hematology and Medical Oncology, Department of Internal Medicine, Faculty of Medicine, Universitas Airlangga, Surabaya 60132, Indonesia
4
Department of Internal Medicine, Airlangga University Hospital, Surabaya 60115, Indonesia
*
Authors to whom correspondence should be addressed.
Hemato 2025, 6(3), 29; https://doi.org/10.3390/hemato6030029
Submission received: 5 April 2025 / Revised: 3 August 2025 / Accepted: 14 August 2025 / Published: 18 August 2025
(This article belongs to the Special Issue Hematopathology: Rare Hematological Diseases)

Abstract

Multiple myeloma (MM) is a malignant plasma cell disorder that evolves from precursor conditions including monoclonal gammopathy of undetermined significance (MGUS) and smoldering multiple myeloma (SMM). Understanding the biological continuum and the molecular drivers of disease progression is crucial for early diagnosis and risk-adapted therapy. Recent advances in next-generation sequencing have identified recurrent mutations in the RAS/MAPK, TP53, and MYC pathways, along with epigenetic alterations that contribute to clonal evolution and therapeutic resistance. Novel diagnostic tools including minimal residual disease (MRD) assessment, gene expression profiling, and advanced imaging have improved risk stratification. Therapeutically, the integration of proteasome inhibitors, immunomodulatory drugs, and anti-CD38 monoclonal antibodies has dramatically improved patient outcomes. In parallel, emerging immunotherapies such as CAR-T cells, bispecific T-cell engagers, and antibody–drug conjugates are expanding treatment options, especially in relapsed or refractory settings. Future directions aim to personalize treatment using genomics, target the tumor microenvironment, and leverage synthetic lethality and epigenetic vulnerabilities. This review highlights the evolving landscape of plasma cell disorders from molecular pathogenesis to cutting-edge therapeutic innovations, emphasizing the need for precision medicine approaches to improve survival and quality of life for patients with MM and its precursors.

1. Introduction

Plasma cell dyscrasias exist on a biologic continuum that ranges from premalignant states to overt malignancy. Monoclonal gammopathy of undetermined significance (MGUS) is the most common of these disorders, affecting over 3% of individuals over the age of 50. It is characterized by the presence of a monoclonal protein (M-protein) in the serum at concentrations <3 g/dL and bone marrow plasma cell infiltration <10%, without symptoms or organ damage [1]. Meanwhile, smoldering multiple myeloma (SMM) is an intermediate stage with higher tumor burden than MGUS, defined by ≥3 g/dL serum M-protein and/or 10–60% bone marrow plasma cells, again without myeloma-defining events [2]. Ultimately, multiple myeloma (MM), the malignant endpoint of this spectrum, is defined by ≥10% clonal bone marrow plasma cells or biopsy-proven plasmacytoma with evidence of end-organ damage (hypercalcemia, renal failure, anemia, or bone lesions) or specific biomarkers of malignancy (e.g., ≥60% bone marrow plasma cells, free light chain ratio ≥100, or >1 focal lesion on MRI) [3]. The risk of progression to symptomatic MM is ~1% per year for MGUS and ~10% per year for SMM [4]. MM accounts for approximately 10% of hematologic malignancies and has a global incidence of ~160,000 cases per year. It remains largely incurable despite therapeutic advancements, with significant morbidity from bone disease, renal failure, and infections [5]. The disease disproportionately affects the elderly and African Americans, and survival varies widely based on molecular risk features and response to therapy. Since MM is almost always preceded by MGUS or SMM, early identification and risk stratification are essential for timely intervention. Advances in molecular diagnostics—such as free light chain assays, gene expression profiling, and imaging—have improved our ability to predict progression [6]. Risk-adapted surveillance and early therapeutic strategies in high-risk SMM have shown promise in delaying or preventing progression to MM [7]. This review aims to provide a comprehensive overview of the molecular and clinical evolution from MGUS and SMM to MM. We examine the role of emerging genomic and epigenetic drivers, risk stratification models, and current therapeutic strategies. Furthermore, we explore the landscape of evolving therapies, including immunotherapies and personalized medicine, with the goal of improving outcomes across the spectrum of plasma cell disorders.

2. Pathophysiology and Disease Continuum

2.1. Clonal Evolution from MGUS to SMM to MM

The progression of MM is widely recognized as a multistep clonal evolutionary process beginning with MGUS, advancing to SMM, and eventually transforming into symptomatic MM (Figure 1). This progression is driven by the accumulation of genetic abnormalities, selective pressures, and interaction with the bone marrow microenvironment. Early genomic events such as hyperdiploidy and immunoglobulin heavy chain (IgH) translocations are frequently observed in MGUS and are thought to initiate clonal plasma cell proliferation. As the disease progresses to SMM and MM, secondary alterations—such as mutations in the RAS/MAPK pathway (KRAS, NRAS, BRAF), MYC dysregulation, TP53 deletions, and chromosomal gains (1q) or deletions (13q, 17p)—confer growth and survival advantages to specific subclones [8]. Importantly, clonal heterogeneity is a hallmark of this process. Subclones with varying genomic profiles coexist, and selective pressures—such as microenvironmental interactions or therapeutic intervention—can promote the dominance of specific aggressive subclones [9]. Two main models of evolution have been proposed (Table 1): linear evolution, where a dominant clone accumulates sequential driver mutations, and branching evolution, where multiple subclones evolve in parallel with distinct mutations and phenotypes [10].
In the linear evolution model (Table 1), a single dominant clone sequentially acquires driver mutations over time, resulting in a unidirectional path toward increasingly aggressive disease. Each new subclone replaces the previous one, creating a dominant “clone of the moment” with a continuous lineage. This model implies a stepwise pattern of progression, often associated with highly aggressive disease and poor prognosis, especially when driven by mutations such as del(17p) or TP53 inactivation [11]. Linear evolution tends to be less influenced by the microenvironment and more by internal genomic instability, leading to rapid clonal sweeps. Patients with linear evolution tend to experience shorter survival post-relapse compared to other evolutionary patterns. In a longitudinal analysis, the median overall survival (OS) for patients with linear evolution was only 25 months, significantly worse than for those with branching or stable patterns [12]. In contrast, the branching evolution model describes a Darwinian diversification process in which multiple subclones evolve in parallel from a common ancestral clone, each acquiring distinct sets of mutations (Table 1). These subclones coexist and compete, often shaped by microenvironmental factors, therapy-induced selective pressures, and immune evasion mechanisms [9]. Branching evolution is the predominant mode observed in MM. In a whole-exome sequencing (WES) study of paired diagnostic and relapse samples, over 70% of patients showed evidence of branching evolution, often with multiple founding clones and fluctuating subclonal dynamics [13]. Importantly, some subclones that are minor at diagnosis may later dominate at relapse, indicating selection under treatment pressure. Branching evolution allows for parallel development of resistance, as different subclones may harbor distinct sensitivities. For instance, one clone may respond to proteasome inhibitors (PIs), while another may resist and expand during therapy, eventually driving relapse [14]. Branching evolution also underlies the emergence of extramedullary disease and minimal residual disease (MRD) escape variants. As such, linear evolution may justify early aggressive treatment to prevent the accumulation of deleterious mutations, whereas branching evolution supports adaptive or combination strategies targeting multiple clones and pathways. Furthermore, branching evolution highlights the need for dynamic risk stratification, as static baseline markers may miss evolving resistance pathways. Clinical trials increasingly aim to incorporate serial molecular assessments to guide therapy selection based on evolving clonal architecture [15].
Furthermore, WES and whole-genome sequencing (WGS) of paired samples from MGUS/SMM and later MM stages demonstrate that many of the driver mutations seen in MM are already present at the precursor stages. For example, NRAS, BRAF, and DIS3 mutations and high-risk cytogenetic abnormalities like t(4;14) and del(17p) may be detected in MGUS or SMM albeit in smaller clonal populations [16]. This suggests that clonal stability rather than rapid mutation accumulation may define many cases of progression, with certain dominant subclones simply expanding over time [17]. Nevertheless, despite the presence of early genomic lesions, disease progression is not solely driven by intrinsic mutations. The bone marrow microenvironment—comprising stromal cells, cytokines, immune components, and extracellular matrix—plays a crucial role in supporting malignant plasma cell survival, immune evasion, and therapy resistance [18]. Additionally, immune dysregulation and loss of immune surveillance, particularly T-cell exhaustion and altered cytokine profiles, contribute to clonal dominance and disease advancement [19]. Thus, the evolution from MGUS to MM is not a sudden transformation but rather a gradual process marked by genetic diversification, subclonal competition, and microenvironmental adaptation.

2.2. Microenvironmental Influences

As mentioned above, the progression from MGUS and SMM to overt MM is not solely dictated by tumor-intrinsic genetic alterations but is profoundly influenced by the bone marrow microenvironment [2]. The bone marrow niche, composed of stromal cells, osteoblasts, osteoclasts, endothelial cells, mesenchymal stem cells (MSCs), immune cells, and the extracellular matrix, provides essential cues that support plasma cell survival, proliferation, and immune evasion (Table 2) [20]. A central mechanism of progression is the remodeling of the stromal and endothelial compartments, which begins during MGUS. MSCs show early transcriptional changes that promote inflammation, while endothelial cells shift toward an angiogenic phenotype in MM, enhancing tumor cell access to nutrients and promoting dissemination [21]. Additionally, disruption of the endosteal niche—marked by increased osteoclast activity and loss of osteoblast function—facilitates tumor proliferation and reduces immune surveillance [22]. The immune microenvironment plays a critical role in allowing malignant plasma cells to escape host defenses. MM cells secrete immunosuppressive cytokines (e.g., TGF-β, IL-10) and interact with regulatory T cells, myeloid-derived suppressor cells, and plasmacytoid dendritic cells to create a tolerogenic and tumor-promoting milieu [23]. Natural killer (NK) cell function becomes increasingly impaired during disease progression, particularly via expansion of dysfunctional CD56bright NK cell subsets with low cytotoxicity [24]. Furthermore, hypoxic conditions in the BM niche induce adaptive responses in MM cells, including metabolic reprogramming, resistance to apoptosis, and enhanced immune evasion through altered antigen presentation and immune checkpoint expression [25]. These adaptations are essential for the persistence of MRD and disease relapse, even after deep therapeutic responses. Collectively, these findings highlight that MM is not a disease solely of rogue plasma cells but also of a corrupted microenvironment that enables tumor survival and immune escape.

2.3. Hallmarks of Malignant Plasma Cell Transformation

The transformation of normal plasma cells into malignant counterparts seen in MM involves a complex interplay of genomic, epigenetic, and microenvironmental alterations (Figure 2). This process aligns with many of the canonical “hallmarks of cancer” and manifests uniquely in the plasma cell lineage. One fundamental hallmark is genomic instability, which drives clonal evolution and progression from MGUS to MM. Malignant plasma cells accumulate chromosomal aberrations—such as hyperdiploidy, IgH translocations (e.g., t(11;14), t(4;14)), and deletions of 13q or 17p—that increase over time and correlate with disease severity. These genetic insults promote cell survival, therapy resistance, and immune evasion [26]. Another critical feature is dysregulated proliferation and evasion of apoptosis, sustained by signaling through IL-6, NF-κB, and PI3K/AKT pathways. Malignant cells upregulate anti-apoptotic proteins such as BCL-2 and MCL-1, fostering survival even under genotoxic stress [27]. Immune evasion is also a hallmark. MM cells reduce antigen presentation, downregulate costimulatory molecules, and recruit immunosuppressive cell populations like regulatory T cells and myeloid-derived suppressor cells. Additionally, hypoxic adaptation in the bone marrow niche induces metabolic and transcriptional shifts that promote resistance to cytotoxic immune cells [2,25]. Epigenetic reprogramming contributes to plasma cell transformation by modifying gene expression independently of DNA sequence. Long non-coding RNAs (lncRNAs) regulate key pathways involved in proliferation, apoptosis, and cell cycle control in MM. Dysregulated expression of lncRNAs, such as MALAT1 and NEAT1, has been implicated in disease progression and drug resistance [28].
MM cells also exhibit a distinct metabolic phenotype characterized by glutamine addiction, a feature that has garnered increasing attention as a potential therapeutic vulnerability. Multiple studies collectively demonstrate that MM cells exhibit a profound dependence on extracellular glutamine (Gln)—a metabolic vulnerability with promising therapeutic implications. Bolzoni et al. showed that MM cells and patient-derived CD138(+) plasma cells produce high levels of ammonium and glutamate in Gln-rich conditions and exhibit high GLS1 (glutaminase) and low GS (glutamine synthetase) expression, confirming their reliance on extracellular Gln. They also identified ASCT2 as the key Gln transporter, where its inhibition significantly reduced MM cell viability in vitro and in vivo [29]. Bajpai et al. further highlighted how Gln withdrawal induces BIM expression and sensitizes MM cells to venetoclax, especially by shifting BIM binding toward BCL-2, suggesting a synthetic lethal strategy involving glutaminase inhibition and BCL-2 blockade [30]. Supporting this, Soncini et al. demonstrated that L-asparaginase (ASNase), a glutaminolytic enzyme, synergizes with carfilzomib by exacerbating mitochondrial stress, ROS generation, and DNA repair disruption, sensitizing even carfilzomib-resistant MM cells [31]. Finally, Effenberger et al. confirmed that glutamine deprivation or GLS1 inhibition with compound 968 induced rapid MYC degradation and apoptosis, especially in MYC-expressing MM cells, while glucose deprivation did not elicit the same apoptotic response [32]. Together, these findings reveal that targeting glutamine metabolism—via transporter blockade, enzymatic depletion, or glutaminase inhibition—may be a potent and selective therapeutic strategy, particularly in combination with venetoclax or PIs.
In addition to metabolic reprogramming, aberrant cellular adhesion and migration are features that enable malignant plasma cells to colonize the bone marrow and form lytic lesions. MM cells display functional diversity in their adhesion properties, with subpopulations exhibiting distinct integrin profiles, migratory behavior, and interaction with extracellular matrix components [33]. Collectively, these hallmarks—genomic instability, apoptotic resistance, immune evasion, epigenetic dysregulation, metabolic reprogramming, and altered adhesion—enable the malignant transformation and expansion of plasma cells, ultimately driving the pathogenesis of multiple myeloma.

3. Genomic and Epigenetic Landscape

3.1. Common Cytogenetic Abnormalities

3.1.1. Hyperdiploidy

Hyperdiploidy, defined as the presence of more than the normal diploid number of chromosomes (typically 47–75), is one of the most common cytogenetic abnormalities in MM, occurring in approximately 45–60% of newly diagnosed patients. It is characterized by trisomies of specific odd-numbered chromosomes, most commonly 3, 5, 7, 9, 11, 15, 19, and 21 [34]. Hyperdiploidy is typically considered an early and initiating event in myelomagenesis, often seen even at the MGUS stage [34]. This cytogenetic pattern defines a distinct molecular subgroup of MM known as “hyperdiploid MM” (H-MM), which is associated with lower rates of IgH translocations and generally more favorable clinical outcomes [35,36]. Several studies have shown that hyperdiploidy confers a survival advantage, particularly in patients lacking high-risk abnormalities such as t(4;14), t(14;16), del(17p), or gain of 1q. For example, in a cohort analysis, hyperdiploid patients had improved PFS compared to non-hyperdiploid counterparts (hazard ratio [HR] 1.9) [37]. However, when hyperdiploidy coexists with high-risk cytogenetic lesions, its protective effect may be mitigated or lost entirely [38]. At the molecular level, H-MM is marked by the upregulation of genes related to protein biosynthesis, ribosomal function, and oxidative phosphorylation, particularly on the gained chromosomes. In a seminal study by Agnelli et al., gene expression profiling of 66 newly diagnosed MM cases revealed that H-MM displayed significant upregulation of 204 genes, the majority of which mapped directly to the gained chromosomes. These genes were functionally linked to translation, transcriptional machinery, and mitochondrial energy production, indicating a gene dosage effect driven by the extra chromosomal content [39]. Further supporting this, a follow-up analysis identified that ribosomal proteins and oxidative phosphorylation components were consistently overexpressed in H-MM, suggesting that this subtype maintains a metabolically active and protein-synthesis-intensive phenotype, potentially reflecting a less aggressive, more differentiated state [40]. These molecular features differentiate it from non-H-MM (NH-MM), which more often involves IgH translocations and cell cycle dysregulation (Table 3). NH-MM shows enrichment for genes involved in cell cycle regulation, proliferation, and oncogene activation (e.g., FGFR3, MMSET, MAF, and MYC), consistent with its more aggressive behavior and poorer prognosis [41]. Moreover, NH-MM exhibits a higher frequency of chromosomal deletions, such as del(13q), del(17p), and del(1p), which are linked to resistance and inferior outcomes. Interestingly, despite their differing genomic origins, both H-MM and NH-MM may converge on shared oncogenic pathways. For example, the deregulation of cyclin D genes—via gene dosage in H-MM and IgH translocations in NH-MM—suggests a unified mechanism of cell cycle dysregulation in MM pathogenesis [42]. Additionally, emerging data suggests that the timing of hyperdiploidy is early, often arising decades before clinical diagnosis. A WGS study revealed that hyperdiploid changes may begin in early adulthood and require years to complete, accumulating secondary mutations during disease latency [43].

3.1.2. IgH Translocations

Translocations involving the IgH locus at chromosome 14q32 are among the most frequent and clinically significant chromosomal alterations in MM. These rearrangements juxtapose oncogenes to the highly active IgH enhancer, leading to deregulated gene expression that drives myelomagenesis. The most common recurrent IgH translocations are t(11;14)(q13;q32), t(4;14)(p16;q32), and t(14;16)(q32;q23), each with distinct biological and clinical implications (Table 4) [2]. The t(11;14) translocation, present in 15–20% of MM patients, involves the cyclin D1 gene (CCND1), promoting cell cycle progression. It is frequently associated with a lymphoplasmacytic morphology and light chain-only subtype. While traditionally considered a standard-risk abnormality, recent evidence reveals clinical heterogeneity depending on coexpressed markers like CD20 and BCL-2 dependency. Importantly, t(11;14)-positive MM exhibits heightened sensitivity to the BCL-2 inhibitor venetoclax, representing the first predictive biomarker in MM therapy [44,45]. The t(4;14) translocation, found in ~15% of MM cases, results in overexpression of FGFR3 and MMSET (NSD2) and is considered high-risk. It is associated with poor response to standard therapies and shortened progression-free survival (PFS) and OS, even in the era of PIs [46,47]. The poor prognosis is thought to be mediated by enhanced oncogenic signaling, epigenetic dysregulation, and chromatin remodeling initiated by MMSET. Next, the t(14;16) translocation is rarer (~4–5%) and leads to overexpression of the MAF oncogene. This lesion is also categorized as high-risk and frequently coexists with additional adverse cytogenetic changes like del(17p) and gain(1q), further worsening prognosis. The pathogenic effects of MAF overexpression include increased cell adhesion, migration, angiogenesis, and deregulated transcriptional programs. Furthermore, t(14;16) MM often harbors an APOBEC mutational signature, contributing to genomic instability and aggressive disease behavior [48,49]. Interestingly, patients with IgH translocations involving undefined partners (t(14;unknown)) have been shown to exhibit favorable survival outcomes in some cohorts, particularly in the context of bortezomib-based regimens, highlighting the diversity and prognostic complexity of IgH rearrangements [50].

3.1.3. Chromosome 13q and 17p Deletions

Deletions of chromosome 13q [del(13q)] are among the most frequently observed chromosomal abnormalities in MM, affecting approximately 50–60% of patients at diagnosis when using sensitive fluorescence in situ hybridization (FISH) techniques [51]. These deletions often involve large segments of the long arm of chromosome 13, typically including the RB1 gene at 13q14, and are frequently associated with monosomy 13. While once considered a uniformly high-risk feature, the prognostic value of del(13q) depends on detection method and co-occurrence with other abnormalities. For instance, deletions detected by metaphase cytogenetics confer a poorer prognosis compared to those detected solely by interphase FISH [52]. Del(13q) is commonly found in combination with other high-risk cytogenetic events, such as t(4;14) and del(17p), and may reflect underlying genomic instability. Moreover, it is more frequently associated with hypodiploid karyotypes and a poor response to standard therapy, especially in patients not treated with PIs [53]. However, novel agents like bortezomib appear to partially overcome the adverse impact of del(13q) [54].
Deletion of chromosome 17p [del(17p)], involving the TP53 tumor suppressor gene at 17p13, is a well-established high-risk marker in MM, present in about 10% of newly diagnosed cases. The prognostic effect of del(17p) is especially pronounced when more than 60–70% of plasma cells harbor the deletion. It is independently associated with shorter PFS and OS, even with high-dose therapy and novel agents [55]. In fact, del(17p) is frequently coexistent with TP53 mutations, and the combined inactivation of both alleles contributes to an aggressive clinical course, marked by drug resistance and rapid relapse [56]. Importantly, the adverse impact of del(17p) cannot be fully mitigated by current treatments, highlighting the urgent need for novel therapeutic approaches, including TP53-independent strategies and enhanced immune-based interventions. Current guidelines recommend integrating del(17p) status into all major risk stratification systems and clinical trial designs.

3.2. Emerging Driver Mutations

3.2.1. RAS Pathway Mutations

Activating mutations in the RAS-RAF-MAPK pathway—specifically in KRAS, NRAS, and BRAF—are among the most frequently observed somatic alterations in MM (Table 5), found in approximately 40–55% of newly diagnosed cases [57]. These mutations contribute to aberrant downstream signaling that promotes tumor cell survival, proliferation, and resistance to therapy. KRAS and NRAS mutations are the most common, affecting codons G12, G13, and Q61. These mutations tend to be mutually exclusive and often occur in subclonal populations, highlighting their role in intratumoral heterogeneity and disease evolution [58]. Notably, NRAS mutations—particularly at codon Q61—have been associated with resistance to PIs like bortezomib, leading to inferior clinical responses and shorter time to progression [59]. BRAF mutations, though less frequent (observed in 4–15% of cases), primarily involve the V600E variant. This mutation activates the mitogen-activated protein kinase (MAPK) pathway independently of upstream signals and has been associated with more aggressive disease and shorter PFS in some cohorts [60]. BRAF-mutant MM cells show increased proteasome activity and reduced ER stress, contributing to drug resistance and survival advantage [61]. Functionally, these mutations converge on the MAPK signaling cascade, promoting phosphorylation of downstream effectors such as MEK and ERK. This cascade enhances proteasome capacity, reduces unfolded protein response activation, and distances malignant cells from apoptosis [61]. Importantly, pharmacologic inhibition of BRAF (e.g., with vemurafenib) or MEK can resensitize MM cells to PIs, especially in those harboring BRAF V600E or activating RAS mutations [62]. Clinically, patients with RAS/MAPK pathway mutations tend to have a higher tumor mutational burden and more rapid progression, particularly in the relapsed/refractory setting [63]. Thus, mutation testing for KRAS, NRAS, and BRAF is becoming increasingly relevant for prognostication and for guiding the use of combination targeted therapies, such as PIs with MEK or BRAF inhibitors.

3.2.2. TP53 and DNA Repair Genes

Mutations and deletions involving the TP53 gene, located on chromosome 17p13, are among the most critical high-risk genetic events in MM. TP53 functions as a tumor suppressor, regulating key cellular processes such as DNA repair, apoptosis, and cell cycle arrest. Deletion of 17p (del(17p)), which encompasses TP53, occurs in approximately 10% of newly diagnosed MM patients, and TP53 mutations often co-occur in this subset, compounding the adverse prognosis [56]. These mutations are most frequently located in the DNA-binding domain and are primarily missense changes that disrupt TP53’s transcriptional activity. Patients with TP53 mutations in the context of del(17p) exhibit significantly shorter PFS and OS, with studies reporting median survival as low as 4 months in high-clonality cases [65]. Gene expression profiling has revealed that low TP53 expression correlates with a unique signature of deregulated genes involved in apoptosis, DNA repair, and cell cycle regulation, forming a distinct high-risk subgroup [66]. In parallel, mutations in DNA damage response (DDR) and DNA repair genes are increasingly recognized as key contributors to MM pathogenesis and resistance to therapy. Alterations in genes involved in homologous recombination (e.g., ATM, BRCA1/2), non-homologous end joining, and base excision repair lead to genomic instability and clonal evolution [67]. Notably, circulating tumor DNA (ctDNA) analyses have shown that DNA repair gene mutations are more prevalent and clonally expanded in relapsed/refractory MM (RRMM) than at diagnosis, with high mutational burden predicting poorer OS [68]. Targeting these DNA repair vulnerabilities is a growing therapeutic focus. For example, CDK12 inhibition with THZ531 suppresses DNA repair gene expression and induces synthetic lethality when combined with PARP (e.g., olaparib) or DNA-PK inhibitors (KU-0060648), offering promising preclinical results in MM models [69].

3.2.3. MYC Deregulation

Deregulation of the MYC oncogene, located on chromosome 8q24, is a pivotal event in the pathogenesis and progression of MM. Unlike primary translocations seen at diagnosis (e.g., IgH translocations), MYC abnormalities typically arise during the transition from SMM to symptomatic disease, making them one of the earliest and most consistent secondary genetic events in MM evolution [70]. Approximately 30–50% of MM cases harbor structural variants involving MYC, including translocations with immunoglobulin enhancers (e.g., IgH, IgK, IgL), amplifications, and complex rearrangements. These events lead to aberrant MYC overexpression and transcriptional activation of downstream targets involved in metabolism, proliferation, and survival [71]. Importantly, the presence of MYC rearrangements correlates with increased tumor burden and poor clinical outcomes, particularly in relapsed or high-risk MM [72]. Beyond structural alterations, recent studies have revealed epigenetic mechanisms driving MYC activation. In cases lacking MYC rearrangements, enhancer elements with increased chromatin accessibility—bound by MM-specific transcription factors such as IRF4 and MAF—can elevate MYC expression via a super-enhancer mechanism. These regulatory elements are enriched in active histone marks (e.g., H3K27ac) and are found to be amplified in a subset of MM patients, contributing to MYC overexpression without genomic rearrangement [73]. Functionally, MYC acts as a transcriptional amplifier, upregulating genes involved in ribosome biogenesis, glucose metabolism, cell cycle progression, and suppression of apoptosis. MYC also cooperates with chromatin regulators and lncRNAs like PVT1, located adjacent to the MYC locus, further stabilizing its oncogenic activity [74]. Importantly, MYC deregulation is tightly associated with disease progression. Its overexpression distinguishes MM from MGUS and SMM and contributes to drug resistance, including reduced sensitivity to immunomodulatory drugs and PIs [75]. Inhibition of MYC activity through direct (e.g., c-Myc/Max disruptors) or indirect approaches (e.g., BET inhibitors like JQ1) has shown efficacy in preclinical MM models by reversing MYC-driven transcriptional programs and inducing tumor regression [76].

3.2.4. Mutational Signatures and Timing in Progression

The transformation from MGUS to MM involves the sequential accumulation of genetic alterations shaped by distinct mutational processes. These processes leave behind characteristic patterns—mutational signatures—that can be used to reconstruct the temporal and mechanistic landscape of disease evolution. Recent WGS studies have identified key mutational signatures that dominate at various phases of MM development. A landmark study analyzing 89 WGS and 973 exomes revealed that many MM-related mutational events are acquired decades before diagnosis, with some key chromosomal gains occurring in early adulthood, suggesting a prolonged subclinical evolution [77]. This early phase is dominated by age-related signatures (SBS1, SBS5) and activation-induced cytidine deaminase (AID) activity (SBS9), indicating their foundational role in initiating clonal plasma cell expansion. As disease progresses toward symptomatic MM, additional mutational processes emerge. In particular, APOBEC-mediated mutagenesis (SBS2 and SBS13) becomes prominent and is associated with disease progression, subclonal diversification, and poor clinical outcomes [78]. APOBEC mutations are especially enriched in RRMM and contribute to genomic instability. This pattern supports a two-phase model: early transformation is driven by AID and age-related mutations, while late progression is fueled by APOBEC activity and other stress-related mutational processes. A broader mutational burden—termed tumor mutational burden (TMB)—also increases significantly from MGUS to MM, particularly in non-synonymous variants and 3′/5′ UTR regions, which may disrupt regulatory mechanisms. High TMB correlates with APOBEC activity and is linked to poorer survival outcomes [79]. Notably, these mutational processes operate in a clonally structured manner. Clonal mutations (early events) are enriched for AID and aging signatures, whereas subclonal mutations (later events) reflect APOBEC and additional novel signatures, such as MM-1 and Signature #8—indicating evolving mutational mechanisms during progression [78]. Finally, longitudinal studies of paired MGUS/SMM and MM samples confirm that most driver mutations and chromosomal changes are already present at the precursor stage. However, their relative clonal abundance changes over time, suggesting that clonal expansion rather than new driver acquisition is often responsible for the transition to symptomatic disease [80].

3.3. Epigenetic Dysregulation

Epigenetic dysregulation is a defining feature of MM pathogenesis and progression [2]. Among the major epigenetic mechanisms, histone modifications and DNA methylation play pivotal roles in regulating chromatin accessibility and gene expression. These changes contribute to both early plasma cell transformation and the development of drug resistance and clonal evolution during disease progression. Histone modifications, such as methylation and acetylation, are catalyzed by histone-modifying enzymes and dynamically control gene transcription through chromatin remodeling. In MM, enzymes such as MMSET (WHSC1), which is overexpressed due to the t(4;14) translocation, are critical in driving tumorigenesis via increased H3K36 methylation, altering chromatin states and transcriptional programs in favor of malignant proliferation [81]. Additionally, UTX, a histone demethylase responsible for removing repressive marks (e.g., H3K27me3), is frequently inactivated in MM via mutations, deletions, or promoter hypermethylation, leading to abnormal gene silencing and worse outcomes [82]. Beyond individual enzymes, global patterns of histone methylation can be detected even in circulating nucleosomes, and levels of H3K9me1—a mark associated with heterochromatin and gene repression—are significantly altered in MM patients’ plasma, suggesting systemic changes in chromatin regulation [83]. DNA methylation, another major epigenetic mark, primarily occurs at CpG dinucleotides and is mediated by DNA methyltransferases (DNMTs). In MM, global hypomethylation coexists with site-specific hypermethylation, particularly at tumor suppressor genes, leading to their silencing. This duality contributes to chromosomal instability and loss of regulatory control over key signaling pathways. For example, DNA methylation of the promoter region of UTX was found to be a frequent mechanism of its inactivation in MM [82]. Importantly, histone modifications and DNA methylation are interconnected. Studies have demonstrated that repressive histone marks like H3K27me3 often precede DNA methylation, reinforcing gene silencing in a coordinated manner [84]. This crosstalk forms a stable epigenetic landscape that supports long-term transcriptional repression in malignant plasma cells [85]. These insights into MM epigenetics have also informed therapeutic innovation. Histone deacetylase inhibitors (HDACis) and DNMT inhibitors (DNMTis) are under clinical evaluation for MM, particularly in combination with PIs and immunomodulatory agents. They aim to reverse aberrant silencing of tumor suppressors and resensitize resistant clones to therapy [81].
Epigenetic regulators such as enhancer of zeste homolog 2 (EZH2), histone deacetylases (HDACs), and bromodomain and extra-terminal domain (BET) proteins play central roles in transcriptional control and chromatin organization in MM, contributing to both disease progression and treatment resistance. The histone methyltransferase EZH2 is the catalytic subunit of the polycomb repressive complex 2 (PRC2), which deposits the repressive histone mark H3K27me3. In MM, EZH2 is frequently overexpressed and correlates with poor prognosis by promoting silencing of tumor suppressor genes and dysregulation of cell cycle checkpoints. Inhibition of EZH2 leads to derepression of cyclin-dependent kinase inhibitors and induction of apoptosis in MM cell lines [86]. Moreover, EZH2 inhibitors have demonstrated potent synergy with HDACis such as panobinostat, producing significant anti-myeloma activity and large-scale transcriptomic remodeling in preclinical models [87]. HDACs regulate chromatin condensation and gene silencing by removing acetyl groups from histone tails. Aberrant HDAC activity has been linked to MM pathogenesis by suppressing genes involved in apoptosis and differentiation. Inhibitors of class I and II HDACs, such as panobinostat and ricolinostat, are under clinical investigation or approved in RRMM settings. HDAC inhibition not only induces apoptosis but also disrupts MM-induced suppression of osteoblast differentiation, particularly through reversal of repressive chromatin at key loci like Runx2 [88]. BET proteins, including BRD4, act as “readers” of acetylated histones, recruiting transcriptional machinery to active chromatin regions. BET proteins regulate key oncogenic drivers, such as MYC and BCL2, and are essential for sustaining MM cell proliferation. BET inhibitors like JQ1 disrupt these transcriptional programs and induce cell cycle arrest and apoptosis in MM cells. However, a notable resistance mechanism involves compensatory upregulation of HDAC6, which diminishes BET inhibitor efficacy. Co-inhibition of HDAC6 and BET proteins has demonstrated strong synergy, reducing MYC expression and enhancing MM cell death in both cell lines and patient-derived samples [89]. Combinatorial approaches targeting these pathways simultaneously—e.g., EZH2 + HDAC inhibitors or BET + HDAC6 inhibitors—represent promising therapeutic strategies. These dual-epigenetic blockade regimens enhance cytotoxicity and reprogram malignant transcriptional networks more effectively than monotherapy [90].
MicroRNAs (miRNAs) are short non-coding RNAs (typically 19–25 nucleotides) that regulate gene expression post-transcriptionally by targeting messenger RNAs for degradation or translational inhibition. In MM, numerous miRNAs are dysregulated, functioning as either oncogenes (oncomiRs) or tumor suppressors, thereby influencing cell proliferation, apoptosis, angiogenesis, drug resistance, and interaction with the bone marrow microenvironment [91]. Tumor-suppressive miRNAs, such as miR-29b, miR-34a, and miR-192/194/215, are frequently downregulated in MM [92,93]. For instance, overexpression of miR-29b in MM cell lines induces apoptosis by directly downregulating the anti-apoptotic protein MCL-1, highlighting its therapeutic potential [94]. Conversely, oncogenic miRNAs such as miR-21 and miR-765 are upregulated in MM and promote tumor cell survival, proliferation, and drug resistance. For example, miR-765 exerts its oncogenic function by targeting and downregulating SOX6, a tumor suppressor gene, thereby enhancing MM cell growth [95]. MiRNAs are also intimately connected with other epigenetic mechanisms in MM. MiR-29 family members target DNMTs, and their loss contributes to global DNA hypermethylation. At the same time, c-MYC suppresses miR-29, creating a feedback loop involving c-MYC, miR-29, DNMTs, and miR-34 family, which is silenced by methylation. Disruption of this loop promotes genomic instability and MM progression [96]. Importantly, miRNAs also contribute to drug resistance and relapse. Dysregulated miRNA profiles in extracellular vesicles (EVs) and exosomes influence the tumor microenvironment and can modulate response to therapy. For instance, specific circulating miRNAs (e.g., miR-140-3p, miR-143-3p) have been correlated with high-risk disease and progression to plasma cell leukemia [97]. Due to their regulatory breadth and disease-specific expression patterns, miRNAs have emerged as attractive biomarkers for diagnosis, prognosis, and monitoring of minimal residual disease. Furthermore, miRNA-based therapies, including mimics for tumor-suppressive miRNAs and inhibitors for oncomiRs, are under preclinical and clinical evaluation as adjunctive strategies in MM [98,99,100].

4. Disease Detection and Risk Stratification

4.1. Diagnostic Criteria

The International Myeloma Working Group (IMWG) provides a tiered diagnostic framework for plasma cell disorders, including MGUS, SMM, and MM (Table 6). These categories represent a biologic continuum, with escalating risk of progression from asymptomatic to symptomatic malignancy. MGUS is defined by the presence of an M-protein <3 g/dL, <10% clonal plasma cells in the bone marrow, and the absence of myeloma-defining events (MDEs) or end-organ damage attributable to plasma cell dyscrasia. MGUS carries a 1% annual risk of progression to MM or related disorders [1]. SMM represents an intermediate state, characterized by M-protein ≥3 g/dL and/or 10–60% clonal bone marrow plasma cells, with no end-organ damage or biomarker-defined events. Risk of progression is variable, averaging 10% per year for the first five years. Risk models such as the “20/2/20” score (BM plasma cells >20%, M-protein >2 g/dL, FLC ratio >20) help stratify progression risk [101]. Traditionally, the diagnostic criteria for MM were based primarily on clonal bone marrow plasma cells ≥10% or a biopsy-proven plasmacytoma and the presence of symptomatic end-organ damage. This was commonly referred to by the acronym CRAB, which is defined as follows: C: calcium elevation—serum calcium >11 mg/dL or >1 mg/dL above the upper limit of normal; R: renal insufficiency—serum creatinine >2 mg/dL or creatinine clearance <40 mL/min; A: anemia—hemoglobin <10 g/dL or >2 g/dL below the normal limit; B: bone lesions—one or more osteolytic lesions on skeletal radiography, CT, or PET-CT. In 2014, the IMWG expanded the diagnostic criteria for MM by introducing biomarker-based myeloma-defining events, collectively known as SLiM-CRAB: S: ≥60% clonal plasma cells in bone marrow; Li: involved–uninvolved free light chain (FLC) ratio ≥100 (with involved FLC ≥100 mg/L); M: >one focal lesion ≥5 mm on MRI; CRAB: classic criteria (hypercalcemia, renal failure, anemia, bone lesions). Patients meeting one or more SLiM or CRAB criteria are now diagnosed with MM and are candidates for therapy initiation [102]. These changes allow for earlier intervention in high-risk patients—those with an ~80% 2-year risk of progression—even in the absence of symptoms. The incorporation of SLiM markers has led to increased MM diagnoses and greater reliance on advanced imaging (MRI, PET/CT), especially for detecting occult bone lesions [103]. Recent meta-analyses suggest that the prognosis of SLiM-positive MM has improved in the modern era, with median time to progression extending to over 30 months in some cohorts [104].

4.2. Molecular and Genomic Risk Markers

Gene expression profiling (GEP) has emerged as a powerful molecular tool to stratify risk in MM by capturing the transcriptional activity of myeloma cells and identifying biologically distinct subgroups. Unlike cytogenetic tests that identify static genetic abnormalities, GEP reflects dynamic gene regulation, making it especially valuable for assessing tumor behavior and treatment response. The most well-established GEP models include the UAMS 70-gene signature (GEP70), IFM-15, and EMC92, all of which can identify high-risk patients who are more likely to relapse early or exhibit resistance to therapy. For instance, in a cohort receiving lenalidomide and dexamethasone, patients classified as high-risk by GEP70 had a median OS of only 19 months compared to unreached survival in low-risk patients, demonstrating its strong predictive power even with novel agents [105]. GEP also enables classification of MM into transcriptional subtypes that correlate with known chromosomal abnormalities (e.g., t(4;14), t(11;14), 1q gain) and clinical phenotypes. High-risk groups frequently show upregulation of genes related to proliferation (e.g., AURKA), centrosome activity, and MYC signaling [106]. In clinical practice, efforts have been made to integrate GEP data with standard risk markers (e.g., International Staging System [ISS] stages, see Table 7; cytogenetics) into unified scoring systems such as the HM-metascore, which combines GEP with chromosomal translocations, proliferation index, and druggable gene expression to produce a composite risk assessment [107]. This approach can delineate three prognostic groups with distinct 6-year survival rates of ~89%, 61%, and 19%. Notably, GEP signatures remain predictive after relapse and can track disease evolution. Serial sampling has shown that transcriptional profiles may shift toward higher-risk patterns over time, often reflecting increased proliferation and genomic instability [108]. Therefore, GEP not only informs initial risk assessment but also has potential utility in monitoring progression and treatment adaptation. Although technical barriers and cost have historically limited its widespread adoption, the development of automated tools and open-source platforms is making GEP more accessible in routine settings. These systems now allow clinicians to interpret GEP data alongside conventional factors in real time to guide therapy choices [109].
Next-generation sequencing (NGS) technologies have transformed the landscape of risk assessment in MM, enabling deep interrogation of the mutational spectrum, clonal architecture, and temporal evolution of the disease [2]. Unlike traditional cytogenetics and gene expression profiling, NGS can simultaneously detect point mutations, structural variants, and copy number abnormalities with high resolution, allowing for personalized prognostication and treatment selection. One of the key applications of NGS is in identifying high-risk genetic features that predispose to early relapse and poor survival. Data from the CoMMpass study—a prospective, longitudinal cohort of newly diagnosed MM patients—show that mutations in TP53, IGLL5, and RAS pathway genes, as well as λ-chain translocations, independently predict early progression and reduced OS [111]. This highlights the importance of integrating mutational profiles into baseline risk stratification models. In parallel, efforts have been made to revise existing clinical tools using NGS. A Revised International Staging System based on NGS (R2-ISS) has been proposed (Table 7), which substitutes FISH with NGS-based detection of high-risk cytogenetic abnormalities (e.g., del(17p), t(4;14), t(14;16)). This NGS-enhanced model showed improved sensitivity for high-risk classification and superior discrimination of event-free survival compared to FISH-based R-ISS [112]. NGS also enables longitudinal tracking of clonal evolution, capturing the emergence of new subclones during therapy and relapse. This dynamic view supports adaptive treatment approaches and early identification of resistance mechanisms [113,114,115]. For example, detection of secondary TP53 mutations or gain of 1q during follow-up is often associated with aggressive disease transformation. Another innovative application of NGS is the integration of multi-omic data—including RNA-seq, WES, and WGS—to build composite risk models. The MMNet-286 gene signature, developed from RNA-seq data in the CoMMpass cohort, effectively stratifies patients by risk of early relapse and correlates with genomic instability and MYC pathway activation [116]. Moreover, NGS is being applied to precursor conditions such as SMM to distinguish high-risk progressors. High-risk SMM patients demonstrate increased mutational load, MAPK pathway activation, and frequent copy number alterations, including 1q gain and 8p loss—features detectable via NGS before clinical progression [117].
Next, MRD detection represents a transformative advancement in the monitoring and risk stratification of MM. MRD refers to the small number of malignant plasma cells that persist in a patient after achieving clinical remission, often undetectable by conventional methods like immunofixation or bone marrow morphology. High-sensitivity MRD assays provide powerful prognostic information and are increasingly used to guide therapy and assess depth of response. IMWG defines MRD negativity as the absence of clonal plasma cells at a sensitivity threshold of at least 1 in 100,000 cells (10−5), achievable via next-generation flow cytometry (NGF) or NGS [118]. NGF-based approaches (e.g., EuroFlow) and NGS platforms (e.g., LymphoSIGHT) have demonstrated comparable sensitivity and are validated for use in clinical trials and routine practice [119]. Achieving MRD negativity is associated with significantly longer PFS and OS, often outperforming traditional response categories such as complete remission (CR). For example, patients who reach sustained MRD negativity over one year have markedly improved long-term outcomes and lower risk of relapse [120]. Importantly, MRD status has proven prognostic value across all treatment phases—including post-induction, post-transplant, and during maintenance therapy [121]. Several techniques are being used for MRD detection, including NGS, NGF, allele-specific oligonucleotide PCR (ASO-PCR), and imaging (PET/CT, MRI) (Table 8). NGS offers high sensitivity (~10−6), high reproducibility, and the ability to monitor clonal evolution over time [122], while NGF allows immediate processing and immunophenotypic characterization but requires fresh samples and standardized panels. ASO-PCR has limited use due to patient-specific primer requirements and lower sensitivity than NGS. Imaging could complement marrow-based MRD by identifying extramedullary disease or focal lesions missed in marrow sampling [123]. Emerging innovations include mass spectrometry-based blood tests for MRD, which may reduce the need for invasive bone marrow biopsies and overcome sampling errors from marrow heterogeneity [124]. Despite its clear prognostic value, standardization, cost, and access remain challenges. There is still debate on how to act on MRD results in clinical practice—whether to escalate therapy for MRD-positive patients or de-escalate for those with sustained MRD negativity. Nevertheless, MRD is increasingly incorporated as a surrogate endpoint in clinical trials and is expected to become central in future response-adapted therapeutic strategies [125].

4.3. Imaging Modalities

Modern imaging modalities are central to diagnosing, staging, and monitoring MM, offering critical insight into bone marrow infiltration, skeletal destruction, and extramedullary disease. The three most widely recommended techniques—positron emission tomography–computed tomography (PET/CT), magnetic resonance imaging (MRI), and whole-body low-dose CT (WBLDCT)—each contribute complementary information (Table 9). PET/CT, particularly with 18F-FDG, provides both metabolic and anatomic data, making it uniquely effective for detecting active disease, evaluating treatment response, and identifying extramedullary involvement. Studies show PET/CT outperforms skeletal survey in identifying osteolytic lesions and has independent prognostic value both at diagnosis and during remission monitoring [126]. PET/CT can also detect MRD and is recommended for solitary bone plasmacytoma and non-secretory myeloma cases where biochemical markers are insufficient. Meanwhile, MRI, especially whole-body MRI (WB-MRI), is considered the most sensitive modality for detecting diffuse and focal marrow infiltration. It is particularly valuable in smoldering MM and solitary plasmacytoma to reveal occult disease not visible on CT or PET/CT [127]. MRI offers superior soft-tissue contrast and can identify spinal cord compression, vertebral fractures, and subtle marrow changes. Newer sequences like diffusion-weighted imaging (DWI) further enhance sensitivity and MR-based assessments of disease burden [128].
WBLDCT has emerged as a preferred first-line imaging technique, replacing conventional skeletal surveys due to its greater sensitivity for osteolytic lesions, faster acquisition time, and better patient tolerability (Table 9) [129,130]. WBLDCT effectively detects subtle cortical bone changes and is now recommended in the initial staging of newly diagnosed MM. Importantly, radiation dose can be optimized to levels equivalent to or lower than skeletal surveys, maintaining image quality while reducing exposure [131]. Recent comparative studies demonstrate that PET/CT and MRI provide complementary roles, with MRI offering higher sensitivity for marrow infiltration and PET/CT delivering better specificity for active disease and response evaluation [132]. The IMWG now endorses the use of at least one of these cross-sectional modalities (PET/CT, MRI, or WBLDCT) in all patients with suspected MM, with PET/CT preferred for treatment monitoring and MRI for early disease detection [133].

5. Current Therapeutic Strategies

5.1. Frontline Therapy

The treatment paradigm for newly diagnosed MM (NDMM) has evolved significantly with the incorporation of triplet and quadruplet regimens into frontline therapy, aimed at achieving deeper responses, delaying progression, and improving survival. The two foundational triplet combinations—bortezomib, lenalidomide, and dexamethasone (VRd) and daratumumab-containing regimens such as Dara-VRd—are now cornerstones of both transplant-eligible and transplant-ineligible patient care (Table 10). The VRd regimen has long been a standard of care, especially following the SWOG S0777 trial, which demonstrated superior PFS and OS for VRd compared to lenalidomide and dexamethasone (Rd) alone [134]. However, the addition of daratumumab, an anti-CD38 monoclonal antibody, has redefined expectations for frontline therapy. In the PERSEUS phase III trial, transplant-eligible patients receiving Dara-VRd showed markedly improved outcomes, with 48-month PFS of 84.3% versus 67.7% with VRd alone. Rates of complete response or better and MRD negativity (10−5) were significantly higher with Dara-VRd (87.9% vs. 70.1% for CR; 75.2% vs. 47.5% for MRD-negative status) [135]. For transplant-ineligible or deferred-transplant candidates, the CEPHEUS study also demonstrated superior efficacy of the quadruplet Dara-VRd over VRd alone. The MRD-negativity rate at 10−5 was 60.9% with Dara-VRd compared to 39.4% with VRd, and the CR or better rate was 81.2% versus 61.6%, respectively. These deep and sustained responses translated into a 43% reduction in risk of progression or death [136]. In a real-world comparative analysis, daratumumab, lenalidomide, and dexamethasone (DRd) showed improved time-to-next-treatment or death (TTNTD) over VRd in transplant-ineligible patients. The median TTNTD was 37.8 months with DRd compared to 18.7 months with VRd, representing a 42% risk reduction [137]. These findings reinforce the efficacy of daratumumab-containing regimens even in less fit populations. Further supporting the advantage of quadruplet induction therapy, a head-to-head comparison between VRd and daratumumab–bortezomib–thalidomide–dexamethasone (D-VTd) in transplant-eligible patients demonstrated superior response depth with D-VTd. The CR or better rate post-transplant was 90.5% for D-VTd versus 68.6% for VRd [138]. Overall, the accumulating evidence from both clinical trials and real-world data supports the adoption of quadruplet regimens—particularly those including daratumumab—as the new standard of care for newly diagnosed multiple myeloma. These regimens offer higher rates of MRD negativity, longer disease control, and deeper responses across patient subgroups.
Autologous stem cell transplantation (ASCT) remains a cornerstone of treatment in transplant-eligible patients with NDMM despite the emergence of potent induction regimens. ASCT consolidates remission after induction therapy and significantly improves PFS and, in many cases, OS. Evidence from multiple randomized clinical trials and real-world studies confirms that ASCT offers superior outcomes compared to chemotherapy alone. For example, in a real-world Uruguayan cohort of 151 patients treated with high-dose melphalan followed by ASCT, the 36-month OS and PFS were 82.4% and 63.8%, respectively, with a median OS of 98 months—supporting its effectiveness in both high- and low-resource settings [139]. ASCT is particularly important for achieving deep, durable responses. In a large retrospective analysis, 15% of patients who underwent ASCT achieved long-term remission (≥8 years PFS), with higher rates of CR and lower incidence of high-risk features compared to others. This suggests a curative potential in a subset of patients [140]. Importantly, the benefits of ASCT persist in the era of novel agents. The Myeloma XI trial demonstrated that ASCT remains effective even in older patients (≥65), when selected based on fitness rather than age alone. Patients aged 65–75 tolerated ASCT well and derived a significant PFS and OS advantage compared to non-transplanted peers, with hazard ratios of 0.41 (PFS) and 0.51 (OS) [141]. Furthermore, ASCT has shown consistent survival benefit across population-based studies. A California registry analysis of 13,494 NDMM patients found that ASCT recipients had a HR of 0.83 for death, with survival benefits extending across age and treatment cohorts. Notably, delayed ASCT (>12 months after diagnosis) was associated with worse outcomes, reinforcing the value of early transplantation [142]. Current guidelines from IMWG continue to support upfront ASCT as the standard of care for eligible patients. Tandem ASCT may be considered for high-risk cytogenetic subtypes or suboptimal response to the first transplant [143].

5.2. Maintenance and Consolidation Therapy

Lenalidomide is the most widely established maintenance therapy following ASCT in patients with NDMM (Table 11). It functions through direct tumoricidal activity, inhibition of angiogenesis, and immune modulation, and it has consistently been shown to prolong PFS and OS. Two pivotal phase III trials—IFM 2005-02 and CALGB 100104—first established the role of lenalidomide in this setting [144,145,146]. Both demonstrated significant PFS improvements, and CALGB 100104 also showed a survival benefit. A meta-analysis pooling these and the GIMEMA trial data confirmed a statistically significant OS benefit from lenalidomide maintenance post-ASCT [147]. Real-world data from the Connect MM registry further validated these findings. Patients receiving lenalidomide-only maintenance had significantly longer PFS (54.5 vs. 30.4 months) and OS (HR 0.45; p = 0.001) compared to those receiving no maintenance. The benefit was seen without a significant increase in healthcare resource utilization, supporting lenalidomide’s feasibility outside clinical trial settings [148]. Lenalidomide also appears to enhance depth of response post-transplant. In a cohort of 139 patients, 38.1% achieved maximal response during maintenance, and 34.3% of initially MRD-positive patients converted to MRD negativity, which correlated with prolonged PFS [149]. These findings support the concept that maintenance therapy not only sustains remission but can deepen response over time, especially in patients with residual disease. The Myeloma XI phase III trial also reinforced lenalidomide’s superiority over observation alone, with median PFS of 39 months vs. 20 months. Importantly, subgroup analyses showed that both standard-risk and high-risk cytogenetic patients benefited, although the OS advantage was more evident in transplant-eligible patients [150]. Concerns regarding toxicity and second primary malignancies (SPMs) have been noted, particularly hematologic adverse events. However, the risk–benefit ratio remains favorable, as the survival benefit generally outweighs the risk of SPMs [147]. Furthermore, studies suggest that lower starting doses (e.g., 5–10 mg/day) are feasible and tolerable in older patients or those with comorbidities [151].
PIs, particularly bortezomib and ixazomib, have emerged as important maintenance options after ASCT, especially in high-risk MM (Table 11). While lenalidomide remains the standard maintenance therapy for most patients, PI-based maintenance is often preferred in those with adverse cytogenetic features or intolerance to immunomodulatory drugs (IMiDs). A systematic review and meta-analysis evaluating bortezomib and ixazomib maintenance post-ASCT found that PI-based strategies significantly improved PFS (HR 0.75, 95% CI 0.67–0.85) and deepened hematologic responses compared to placebo or thalidomide, without increasing SPM risk [152]. While OS data were less conclusive, these results support the efficacy of PIs in prolonging disease control. Real-world and retrospective data further support the preferential use of PIs in high-risk cytogenetic subgroups, such as patients with del(17p), t(4;14), or gain(1q). A Mayo Clinic cohort showed that high-risk patients on PI-based maintenance had better PFS compared to those on IMiD-only or no maintenance, with the benefit being most pronounced in those who underwent early transplant (<12 months from diagnosis). In this group, bortezomib-containing maintenance improved long-term outcomes and narrowed the survival gap between high-risk and standard-risk patients [153]. In the HOVON-65/GMMG-HD4 trial, bortezomib maintenance for two years post-ASCT showed clear PFS and OS advantages over thalidomide, particularly among patients with high-risk cytogenetics. Notably, the trial also demonstrated a reduction in renal failure, further supporting PI use in renally impaired MM patients [154,155]. Among 827 patients, those receiving bortezomib-based induction (PAD: bortezomib, doxorubicin, dexamethasone) followed by bortezomib maintenance showed significantly higher complete response rates (31% vs. 15% after induction; 49% vs. 34% during maintenance) compared to the VAD (vincristine-based) regimen. After a median follow-up of 41 months, the PAD group had improved PFS (35 vs. 28 months; HR 0.75, p = 0.002) and OS (HR 0.77, p = 0.049). Notably, high-risk subgroups—including patients with elevated creatinine (>2 mg/dL) and those with del(17p13)—experienced substantial survival benefits from bortezomib. For instance, in patients with renal impairment, bortezomib extended PFS from 13 to 30 months and OS from 21 to 54 months [156]. After a median follow-up of 96 months, bortezomib-based treatment still significantly improved PFS (HR = 0.76, p = 0.001), though OS was not statistically different between arms (HR = 0.89, p = 0.24). Importantly, the adverse impact of high-risk features such as deletion 17p13 and renal impairment was mitigated in the PAD arm but persisted in the VAD group. Rates of SPM were similar across arms (7%, p = 0.73), and OS after first relapse was comparable (HR = 1.02, p = 0.85). The study confirms that bortezomib during induction and maintenance provides sustained PFS benefit in MM without increasing long-term toxicity risk [157].
More recently, ixazomib, an oral PI, has shown promise in the TOURMALINE-MM3 and MM4 trials (Table 11). In the TOURMALINE-MM4 study involving transplant-ineligible patients, ixazomib maintenance significantly improved PFS (17.4 vs. 9.4 months) compared to placebo, with a 34% risk reduction in disease progression or death and a favorable safety profile [158]. These findings suggest that ixazomib may be a practical option for patients seeking an oral, less toxic maintenance alternative. Despite these benefits, some limitations remain. Toxicity—particularly peripheral neuropathy with bortezomib—can affect adherence, although weekly dosing and subcutaneous administration have mitigated these effects. Additionally, head-to-head comparisons between lenalidomide and PIs for maintenance are lacking, and ongoing trials aim to clarify optimal maintenance selection and duration.

5.3. Management of High-Risk Disease

High-risk multiple myeloma (HRMM) is defined by the presence of specific cytogenetic abnormalities associated with poor clinical outcomes. Among these, TP53 alterations, especially double-hit MM (coexistence of TP53 deletion and mutation), represent some of the most aggressive and treatment-resistant forms of the disease. A landmark study using SNP array and next-generation sequencing found that biallelic TP53 inactivation (double-hit) conferred the worst clinical outcomes, with significantly shorter PFS and OS compared to single-hit or TP53 wild-type patients. Double-hit MM patients had a PFS HR of 3.34 and an OS HR of 3.47, underscoring their aggressive biology [159]. Another large-scale study confirmed that del(17p) alone also confers poor prognosis, but the combination with TP53 mutation defines the most aggressive subgroup—double-hit MM—with median OS of just 36 months [160]. Ultra-deep sequencing techniques have revealed that TP53 mutations are more prevalent than previously recognized, particularly at relapse, and double-hit MM is associated with younger age at diagnosis, higher marrow plasma cell infiltration, and a higher rate of extramedullary disease [161]. From a therapeutic standpoint, TP53 inactivation confers resistance to alkylating agents like melphalan and doxorubicin, while PIs retain efficacy, even in TP53-deficient models [162]. This observation supports the use of PI-based maintenance (e.g., bortezomib) in patients with del(17p) or TP53 mutation. Double-hit MM is also associated with higher clonal fitness and therapy resistance. Functional in vitro models show that both monoallelic and biallelic TP53 lesions increase resistance to genotoxic agents and enhance cell fitness, especially under drug selection pressure [163]. These findings highlight the urgent need for personalized approaches, including early ASCT and PI-based consolidation in frontline therapy, triplet/quadruplet regimens incorporating daratumumab, and participation in clinical trials evaluating CAR-T, bispecific antibodies, or TP53-reactivating agents. Despite aggressive therapy, outcomes remain suboptimal. A real-world study of double- and triple-hit MM reported median OS of just 13 months, with early deaths from progressive disease common, emphasizing the unmet need in this population [164].
Thus, the management of HRMM—characterized by adverse cytogenetic profiles like del(17p), t(4;14), t(14;16), gain(1q), and TP53 mutations—requires an individualized and intensified therapeutic approach due to the aggressive nature and poor prognosis of the disease (Table 12). Traditional uniform strategies are no longer adequate, and mounting evidence supports risk-adapted regimens designed to overcome resistance and deepen remission. Recent pooled analyses of phase III trials have demonstrated that continuous therapy incorporating PIs and IMiDs significantly improves outcomes in high-risk patients. Specifically, high-risk patients treated with continuous bortezomib-IMiD maintenance had greater PFS and OS compared to those receiving fixed-duration therapy, effectively narrowing the survival gap with standard-risk patients [165]. Quadruplet induction regimens, such as Dara-VRd, are strongly recommended for HRMM due to their ability to achieve deeper and more durable responses, including MRD negativity. In transplant-eligible patients, this approach enhances MRD-negative rates and has been associated with longer PFS compared to triplet regimens [166]. In the post-transplant setting, double or tandem ASCT is a viable strategy for patients with high-risk features. Retrospective analyses show that tandem transplants can lead to deeper responses and may improve survival outcomes, especially in patients with t(4;14) or del(17p), although this benefit is not uniform and needs further validation through prospective trials [167]. Maintenance therapy also needs to be intensified. While lenalidomide is the standard for standard-risk MM, high-risk patients benefit more from combination maintenance with a PI (e.g., bortezomib or carfilzomib) plus lenalidomide, which has shown improved depth of response and survival in high-risk cohorts [168]. Moreover, risk-adapted therapy must be dynamic, guided by MRD status and evolving molecular markers. MRD negativity, sustained over time, is emerging as the most powerful prognostic tool and may eventually guide de-escalation or intensification of therapy [167]. To summarize, tailored intensive strategies for HRMM integrate quadruplet induction, early transplant (often tandem), and dual-agent maintenance, with therapy modulation based on MRD status. These approaches are essential for improving outcomes in a subgroup traditionally associated with rapid relapse and inferior survival.

5.4. Management of Bone Loss in Multiple Myeloma

Bisphosphonates and denosumab are cornerstone agents in the management of bone disease in MM. Both classes of drugs act by inhibiting osteoclast-mediated bone resorption, though through distinct mechanisms—bisphosphonates incorporate into the bone matrix and disrupt osteoclast function upon resorption, while denosumab is a monoclonal antibody that neutralizes RANKL, a key driver of osteoclastogenesis. Bisphosphonates, particularly zoledronic acid and clodronate, have long been used to reduce skeletal-related events (SREs) and bone pain in MM. Zoledronic acid has been shown to significantly reduce the risk of SREs and may also improve OS, although it is associated with nephrotoxicity and a risk of osteonecrosis of the jaw (ONJ) with long-term use [169]. ONJ remains a major concern, occurring in 6–9% of patients, particularly with prolonged therapy. Preventative dental assessments are essential before and during treatment [170]. Meanwhile, denosumab has emerged as a potent alternative, particularly in patients with renal dysfunction. In the pivotal phase III study comparing denosumab with zoledronic acid, denosumab was non-inferior in delaying time to first SRE and demonstrated a more favorable renal safety profile [171]. A systematic review concluded that denosumab offers comparable efficacy with potentially lower renal toxicity, although both agents have similar rates of ONJ [169]. Denosumab is administered subcutaneously and does not require renal dose adjustments, enhancing its convenience in clinical settings. Both drugs are also used in the treatment of hypercalcemia of malignancy, with denosumab being effective even in patients refractory to bisphosphonates [171]. However, hypocalcemia is more common with denosumab, especially in patients with impaired renal function, necessitating calcium and vitamin D supplementation [172].

6. Immunotherapy and Targeted Agents

6.1. Monoclonal Antibodies

Targeting CD38, a transmembrane glycoprotein highly expressed on malignant plasma cells, has revolutionized MM therapy. Two anti-CD38 monoclonal antibodies—daratumumab and isatuximab—are now widely used across treatment lines, from newly diagnosed to RRMM, due to their deep and durable anti-myeloma effects (Table 13). Daratumumab is a fully human IgG1κ monoclonal antibody that induces myeloma cell death through multiple mechanisms, including complement-dependent cytotoxicity (CDC), antibody-dependent cellular cytotoxicity (ADCC), phagocytosis, and direct apoptosis. In addition, it depletes CD38+ immunosuppressive cells, restoring immune effector function [173]. Daratumumab is approved for both transplant-eligible and ineligible newly diagnosed MM patients in combination regimens such as Dara-VRd, which has shown high rates of MRD negativity and prolonged PFS [135]. Isatuximab, while also targeting CD38, binds to a distinct, non-overlapping epitope compared to daratumumab. It exhibits unique properties, including direct pro-apoptotic activity without crosslinking and less reliance on complement activation, making it effective even in low-CD38-expressing tumors [174]. Isatuximab is approved in combination with pomalidomide and dexamethasone (Isa-Pd) or carfilzomib and dexamethasone (Isa-Kd) for patients with RRMM, including those with high-risk features. Real-world evidence supports the continued efficacy of anti-CD38 therapy across multiple lines of treatment, even after prior exposure. In a cohort of relapsed MM patients receiving more than one anti-CD38 regimen, the overall response rate remained stable (~40%) in the first two exposures, suggesting clinical benefit from retreatment strategies [175]. Despite their success, resistance and CD38 downregulation can occur with repeated exposure. Recent studies report biallelic loss of CD38 in a subset of patients post-treatment, conferring resistance to both agents. Interestingly, some CD38 mutations selectively affect daratumumab binding but retain sensitivity to isatuximab, suggesting a potential role for sequential anti-CD38 therapy guided by molecular profiling [176]. Combination approaches to enhance anti-CD38 efficacy are under investigation. Inhibitors of CD46 and CD59, which block CDC, have been shown to increase the cytotoxicity of daratumumab and isatuximab in vitro and in vivo [177].
Next is elotuzumab, a first-in-class, humanized IgG1 monoclonal antibody that targets Signaling Lymphocytic Activation Molecule Family member 7 (SLAMF7), a glycoprotein highly and selectively expressed on both normal and malignant plasma cells as well as NK cells (Table 13). Unlike other monoclonal antibodies, elotuzumab exhibits little direct cytotoxicity as a monotherapy but shows potent activity when combined with IMiDs through immune-cell-mediated mechanisms. Elotuzumab exerts its anti-myeloma effects primarily by enhancing NK cell-mediated ADCC. It binds SLAMF7 on MM cells and simultaneously activates NK cells via CD16, thereby bridging immune effectors with tumor targets [178]. Additionally, elotuzumab can provide costimulatory signals to NK cells via direct SLAMF7 engagement, further enhancing their cytotoxic potential even in CD16-independent pathways [179]. The clinical efficacy of elotuzumab was demonstrated in the ELOQUENT-2 trial, which led to its FDA approval in 2015. In this phase III trial, patients with RRMM who received elotuzumab in combination with lenalidomide and dexamethasone (ELd) had a 30% reduction in the risk of progression or death compared to those receiving lenalidomide and dexamethasone alone [180]. Follow-up studies confirmed the durable benefit and favorable safety profile of ELd in various patient populations, including those with high-risk cytogenetics like t(4;14) and del(17p) [181]. Preclinical and clinical studies also show that elotuzumab synergizes with IMiDs like lenalidomide by promoting NK cell activation through cytokine production (e.g., IL-2, TNF-α) and upregulation of adhesion molecules [182]. This synergy underlines its limited single-agent activity but robust efficacy in combination regimens. While elotuzumab is less frequently used in the frontline setting due to the dominance of anti-CD38 antibodies, it retains value in RRMM, especially for patients previously exposed to daratumumab. Importantly, SLAMF7 expression is retained even in advanced or extramedullary disease, supporting its continued utility in resistant cases [183].

6.2. Antibody–Drug Conjugates

Belantamab mafodotin is a first-in-class antibody–drug conjugate (ADC) targeting B-cell maturation antigen (BCMA), a transmembrane protein highly expressed on malignant plasma cells. It comprises an afucosylated anti-BCMA monoclonal antibody conjugated to the cytotoxic agent monomethyl auristatin F (MMAF). Upon binding to BCMA, the drug is internalized and delivers MMAF directly into the cell, disrupting microtubule dynamics and inducing apoptosis [184]. Belantamab mafodotin was granted accelerated FDA approval in 2020 for patients with RRMM who had received at least four prior therapies, including a proteasome inhibitor, an immunomodulatory drug, and an anti-CD38 antibody. The approval was based on results from the DREAMM-2 trial, which demonstrated a 32–35% overall response rate (ORR) and a median duration of response of 6.2–12.5 months in heavily pretreated patients [185,186,187]. Notably, responses occurred across high-risk cytogenetic subgroups, including patients with extramedullary disease. Despite its clinical activity, ocular toxicity—particularly keratopathy—is a major limiting factor [187]. Up to 69% of patients develop corneal microcyst-like epithelial changes (MECs), often leading to visual disturbances, dose delays, or discontinuation [188]. This adverse event profile prompted the withdrawal of belantamab mafodotin from the U.S. market in 2022 pending confirmatory trial outcomes, though it remains available under compassionate use and in clinical studies.
Ongoing trials are exploring combination strategies to enhance efficacy and reduce toxicity. In the BELARD study, combining belantamab with lenalidomide and dexamethasone (belantamab-Rd) in transplant-ineligible NDMM patients yielded encouraging response rates and a manageable safety profile, suggesting potential synergy with ImiDs [189]. Likewise, in the BELACARD study, belantamab combined with carfilzomib and dexamethasone showed promising efficacy in RRMM, supporting continued development in multidrug regimens [190]. Furthermore, gamma-secretase inhibitors (GSIs; e.g., nirogacestat) that prevent BCMA shedding have demonstrated synergistic activity with belantamab mafodotin in preclinical models, leading to enhanced ADCC and tumor cell killing—even in low-BCMA-expressing cells [191]. This approach is being tested in the ongoing DREAMM-5 [192].

6.3. CAR-T-Cell Therapy

Chimeric antigen receptor T-cell (CAR-T) therapy targeting BCMA has transformed the treatment landscape for RRMM. BCMA is selectively expressed on malignant plasma cells, making it an ideal target for cellular immunotherapy. Two FDA-approved BCMA-targeting CAR-T therapies—idecabtagene vicleucel (ide-cel) and ciltacabtagene autoleucel (cilta-cel)—have demonstrated unprecedented efficacy in heavily pretreated patients (Table 14). Ide-cel was the first BCMA CAR-T product to receive regulatory approval. In the pivotal KarMMa trial, ide-cel achieved an ORR of 73% and median PFS of 8.8 months, with durable responses in patients with a median of six prior therapies [193]. While ide-cel showed manageable rates of cytokine release syndrome (CRS) and neurotoxicity, long-term remissions were limited in many patients due to CAR-T-cell exhaustion and antigen escape. Meanwhile, cilta-cel, a second-generation CAR-T with two BCMA-binding domains, has demonstrated deeper and more durable responses. In the CARTITUDE-1 study, cilta-cel yielded a 98% ORR, with 83% achieving stringent CR (sCR) and median PFS exceeding 34.9 months, setting a new benchmark in RRMM [194]. Subsequent trials, such as CARTITUDE-4, have shown efficacy even earlier in treatment lines, suggesting potential for curative intent [195]. Despite the promise of BCMA CAR-T, relapse remains a concern, occurring in 27–64% of responders. Strategies to improve outcomes include retreatment with alternative BCMA CAR-T constructs and combination with bispecific antibodies targeting other antigens (e.g., FcRH5) [196]. Cilta-cel retreatment has shown feasibility, with 80% response rates in patients previously exposed to ide-cel.
Both ide-cel and cilta-cel are associated with immune-mediated toxicities, particularly CRS and immune effector cell-associated neurotoxicity syndrome (ICANS) (Table 14). These toxicities result from excessive immune activation and inflammation following T-cell engagement and expansion and require vigilant monitoring and prompt management to ensure patient safety. CRS is the most common adverse event following CAR-T therapy, occurring in up to 90% of patients, but is mostly Grade 1–2. It is triggered by massive cytokine release from activated CAR-T cells and myeloid cells, leading to fever, hypotension, hypoxia, and in severe cases, multi-organ dysfunction. The standard treatment for moderate to severe CRS (Grade ≥2) includes tocilizumab, an IL-6 receptor antagonist, and corticosteroids for more severe or refractory cases [197]. Early use of these agents has been shown to reduce toxicity without impairing CAR-T efficacy [198]. In contrast, ICANS is a distinct neurotoxicity syndrome often following CRS, characterized by symptoms ranging from mild confusion and aphasia to seizures and cerebral edema. The pathophysiology involves disruption of the blood–brain barrier and neuroinflammation driven by systemic cytokines. Management of ICANS includes high-dose corticosteroids (e.g., dexamethasone), and unlike CRS, tocilizumab is not effective due to limited blood–brain barrier penetration [199]. For patients with steroid-refractory ICANS, other agents such as anakinra (IL-1R antagonist) and anti-TNF therapies are under investigation [200]. Neurotoxicities, including Parkinsonism-like syndromes and cranial nerve palsies, are rare but more frequently reported with cilta-cel due to its higher expansion kinetics [201]. High peak CAR-T expansion is associated with both improved response and increased neurotoxicity risk [202]. Infections are also a notable complication, with hypogammaglobulinemia and prolonged cytopenias contributing to bacterial and viral infections. However, comparative real-world analyses show similar infection risks between ide-cel and cilta-cel, reinforcing the need for infection prophylaxis and monitoring [203].
The American Society for Transplantation and Cellular Therapy (ASTCT) provides widely accepted grading criteria for CRS (Table 15) and ICANS (Table 16), guiding when to escalate therapy [204]. For example, Grade 1 CRS is managed supportively, while Grade 2 or higher typically warrants tocilizumab and possibly corticosteroids. ICANS is graded based on the Immune Effector Cell-Associated Encephalopathy (ICE) score and neurological findings, with steroids initiated for Grade ≥2 [205]. Timing and duration of these toxicities are critical for post-infusion monitoring. A multicenter study found that late-onset CRS or ICANS beyond day 14 is rare, prompting discussion of shortening the standard 4-week monitoring period to reduce burden on patients and healthcare systems [206]. Importantly, while CRS and ICANS are not entirely avoidable, recent evidence suggests that effective toxicity management does not compromise efficacy. In fact, early intervention may even improve response rates by minimizing immune exhaustion [207].
The resistance to CAR T-cell therapy in MM is driven by a complex interplay of host-related factors, tumor-intrinsic mechanisms, the immunosuppressive tumor microenvironment, and characteristics of the CAR T-cell product itself. While traditional prognostic factors like age are less predictive in CAR T-cell therapy, physical fitness and performance status remain critical, with conditions like sarcopenia and cachexia linked to poor outcomes. Intriguingly, visceral adiposity has a paradoxical role—associated with more severe CRS due to increased IL-6 production from adipose tissue but also with better PFS and OS, possibly through “metaflammation,” a pro-inflammatory transformation of fat tissue that supports immune activation. Additionally, renal function and weight significantly affect fludarabine pharmacokinetics, influencing lymphodepletion efficiency. Notably, broad-spectrum antibiotics prior to infusion were associated with reduced CAR T efficacy and increased neurotoxicity, suggesting the gut microbiome’s critical role in modulating CAR T-cell response [208]. Meanwhile, a key tumor-intrinsic resistance mechanism is antigen loss or downregulation, particularly BCMA. Although biallelic BCMA deletions are rare, partial or transient downregulation without genetic mutations can result in antigen-low escape. This may reflect clonal selection under continuous immunologic pressure. Notably, responses to CAR T-cell therapy have still been observed in patients with prior anti-BCMA therapy exposure, suggesting some capacity to overcome BCMA-low clones. Additional resistance mechanisms include upregulation of inhibitory ligands (e.g., PD-L1), loss of costimulatory molecules (e.g., CD58), and altered apoptotic machinery, which together impair CAR T-cell cytotoxicity even when the target antigen is present [208]. The bone marrow tumor microenvironment in MM also acts as a formidable barrier to CAR T-cell efficacy. It harbors suppressive immune cells such as M2 macrophages, Tregs, and MDSCs, all of which foster CAR T-cell exhaustion and reduce infiltration. For example, STAT3-activated tumor-associated macrophages promote plasma cell growth and resistance to immune killing. Additionally, extracellular matrix components and abnormal vasculature physically restrict CAR T-cell trafficking. Immune profiling studies have also shown that pre-existing inflammatory signatures and serum markers like CRP, IL-6, and ferritin correlate with poor CAR T expansion and response, emphasizing the predictive value of systemic inflammation [208].
Next, the phenotype and fitness of T-cells used in manufacturing also critically affect outcomes. Products enriched for less differentiated (naïve or early memory) T cells, especially CCR7+CD45RA+ CD8+ T-cells, exhibit better expansion and persistence, translating into superior efficacy. Conversely, T-cell exhaustion, driven by chronic antigen stimulation and tonic CAR signaling, leads to functional decline, marked by upregulation of PD-1, LAG-3, TIM-3, and TIGIT. This is accompanied by epigenetic reprogramming and loss of cytokine production. Moreover, some patients exhibit CAR-expressing regulatory T cells (CAR-Tregs) post-infusion, which can suppress overall CAR T expansion and function, further compromising response [208]. Finally, a critical but underappreciated barrier to durable CAR T efficacy is immune rejection of the CAR T-cells themselves. Most CAR constructs use murine-derived single-chain variable fragments (scFvs), which are immunogenic. Patients may develop anti-CAR antibodies or CAR-specific cytotoxic T cells that eliminate infused cells. Such responses are associated with poor CAR T-cell expansion and relapse after initial remission. Evidence from MM trials using biepitopic CARs showed anti-CAR antibody development post-relapse, often coinciding with CAR T-cell clearance. Immune responses can arise from both humoral mechanisms and cross-presentation of murine peptides by antigen-presenting cells, which prime endogenous T cells to target the CAR [208].

6.4. Bispecific T-Cell Engagers (BiTEs)

BiTEs represent a new frontier in immunotherapy for RRMM, utilizing engineered antibodies to simultaneously bind a tumor antigen and CD3 on T cells, thereby redirecting cytotoxic T cells to eliminate malignant plasma cells (Figure 3). The most validated BiTE targets in MM are BCMA and the more recently identified G-protein-coupled receptor class C group 5 member D (GPRC5D). BCMA × CD3 BiTEs, such as teclistamab, were among the first to reach clinical application. In heavily pretreated RRMM patients, teclistamab demonstrated response rates of approximately 60% in early-phase trials, including high-risk subgroups [209]. These agents activate T cells independent of costimulatory signals and are highly potent even in patients previously treated with CAR-T or anti-CD38 antibodies [210]. Major adverse events include CRS and cytopenias, which are generally manageable with corticosteroids and IL-6 inhibitors. However, relapse after BCMA-targeted therapies is often driven by antigen escape or downregulation, prompting development of agents targeting GPRC5D, an antigen broadly expressed on MM cells but limited in normal tissues. Talquetamab, a GPRC5D × CD3 BiTE, has shown robust preclinical and clinical efficacy. Preclinical studies using constructs such as BR109 demonstrated potent, selective T-cell-mediated cytotoxicity in vitro and in xenograft mouse models, validating GPRC5D as a promising alternative to BCMA [211]. Moreover, GPRC5D-targeting BiTEs may overcome resistance seen in BCMA-exposed patients. Data indicate GPRC5D is more consistently and homogeneously expressed than BCMA, even in post-CAR-T or BiTE relapse settings [212]. Clinical development is ongoing for several GPRC5D × CD3 agents, including forimtamig and MBS314, which also show potential in targeting dual-negative clones [213,214]. To combat tumor heterogeneity and antigen escape, next-generation constructs are exploring trispecific formats, combining BCMA, GPRC5D, and CD3 arms. Agents like JNJ-79635322 and GBD218 have shown the ability to target dual-positive and single-positive cells, enhance T-cell engagement, and reduce cytokine release, all while retaining potent in vivo antitumor activity [215,216].

7. Future Directions in Therapeutics

7.1. Targeting Other Mutational Pathways

MAPK pathway, particularly the RAS/RAF/MEK/ERK axis, plays a central role in the regulation of cellular proliferation, differentiation, and survival. Mutations that activate this pathway are highly prevalent in MM, with NRAS, KRAS, and BRAF mutations collectively found in up to 50% of newly diagnosed MM cases. These alterations lead to constitutive signaling and are strongly associated with disease progression, drug resistance, and poor prognosis [217]. Given its role as a key oncogenic driver, the MAPK pathway has become a promising therapeutic target. MEK inhibitors, such as trametinib and selumetinib, have demonstrated cytotoxic effects against MM cells, especially those harboring activating RAS or BRAF mutations. MEK inhibition alone, however, often produces only modest effects due to compensatory signaling through parallel survival pathways [218]. A more effective strategy may involve combination therapy targeting both MAPK and PI3K/AKT pathways, which frequently co-activate in MM. In primary MM samples, dual inhibition of MEK/MAPK and PI3K/AKT signaling resulted in synergistic cytotoxicity in over 75% of cases, particularly those with RAS mutations [219]. In a biomarker-driven phase II clinical trial, trametinib was tested in RRMM patients stratified by RAS/RAF mutation status. Among 25 enrolled patients, the ORR to trametinib monotherapy was only 8%, with 17% achieving clinical benefit. However, when combined with the AKT inhibitor GSK2141795, the ORR improved to 27%, suggesting that co-targeting the MAPK and PI3K/AKT pathways may disrupt feedback loops and circumvent adaptive resistance mechanisms [220]. In the trial, thrombocytopenia and gastrointestinal symptoms were among the most common adverse events, but overall tolerability was acceptable. Notably, while the study did not demonstrate strong single-agent efficacy, it provided proof-of-concept that dual pathway blockade can enhance cytotoxicity in MM [220].
Targeting the p38 MAPK subpathway, which modulates the tumor microenvironment and cytokine signaling, is another emerging strategy. Inhibition of p38α MAPK using agents like VX-745 and SD-282 significantly reduced MM cell proliferation, IL-6 secretion, and tumor angiogenesis, particularly in the bone marrow niche [221,222]. More recently, dual inhibition strategies targeting MEK and MAP4K2—kinase-regulating stress-activated MAPK branches—have shown potent synergistic effects in RAS-mutated MM cells, including profound suppression of IKZF1/c-MYC and enhanced apoptosis [223]. This approach may provide a framework for personalized therapy in genomically defined subgroups. Furthermore, crosstalk between the MAPK pathway and DDR suggests that co-targeting these networks may enhance genomic stress and sensitize MM cells to apoptosis, especially in relapsed/refractory disease [224].
Next, synthetic lethality is a promising therapeutic approach that exploits the unique vulnerabilities of cancer cells harboring DNA repair defects. In MM, where genomic instability and defective homologous recombination (HR) repair are common, targeting compensatory DNA repair pathways—particularly with poly (ADP-ribose) polymerase (PARP) inhibitors—has gained momentum. MM cells often exhibit intrinsic chromosomal instability and dependence on HR to maintain survival, especially given frequent deficiencies in the non-homologous end-joining (NHEJ) pathway. This makes them uniquely susceptible to synthetic lethality strategies. PARP inhibitors, such as olaparib and veliparib, prevent repair of single-strand breaks, leading to accumulation of lethal double-strand breaks during DNA replication, particularly when HR is impaired [225]. Pharmacologic induction of “BRCAness” is another novel approach being explored. CDK inhibitors such as dinaciclib and THZ531 downregulate BRCA1/2 and RAD51 expression, mimicking HR deficiency and sensitizing MM cells to PARP inhibitors. In vitro and in vivo studies combining CDK inhibitors with PARP inhibitors (e.g., olaparib or ABT-888) have demonstrated synergistic cytotoxicity and tumor regression while sparing normal B cells [69,225]. In addition to PARP, ATR inhibition represents another synthetic lethality strategy. ATR is essential for responding to replication stress and DNA cross-links. In preclinical models, combining ATR inhibitors (e.g., VX-970) with melphalan, a frontline MM alkylator, led to synergistic anti-myeloma effects, even in melphalan-resistant lines. This strategy takes advantage of MM cells’ addiction to ATR signaling and has been shown to significantly prolong survival in xenograft models [226]. Moreover, combination therapies of PARP inhibitors with PIs like bortezomib have shown cell-line-specific synergy. In certain MM models, veliparib enhanced bortezomib cytotoxicity by reducing its effective dose sevenfold—suggesting that DNA damage responses modulate proteasome stress pathways [227]. These synthetic lethality strategies are especially attractive for HRMM or RRMM patients who harbor HR repair gene mutations (e.g., TP53, ATM, BRCA1/2) or show treatment resistance. Ongoing trials are needed to determine optimal biomarker-driven patient selection, dosing, and safety in combination with immunotherapy or cytotoxic agents.

7.2. Epigenetic Modulators

Epigenetic dysregulation is a hallmark of MM progression and therapeutic resistance. Abnormal expression of chromatin modifiers—such as BET proteins, HDACs, and EZH2—has been implicated in maintaining oncogenic transcriptional programs, particularly through upregulation of MYC, BCL-2, and other survival pathways. Thus, targeting these epigenetic regulators represents a promising approach for overcoming drug resistance and improving disease control in MM. BET inhibitors (e.g., JQ1, BAY 1238097) suppress transcription by preventing BET protein (such as BRD4) binding to acetylated histones. This leads to downregulation of oncogenes like MYC, BCL-xL, and cyclin D1. BET inhibition alone reduces MM cell proliferation and induces apoptosis, and it has shown synergistic effects when combined with other epigenetic agents [228]. However, resistance can arise through compensatory HDAC6 upregulation, suggesting the need for rational combination therapies [89]. HDAC inhibitors, such as panobinostat and vorinostat, block histone deacetylation, restoring transcriptional activation of tumor suppressor genes and triggering apoptosis. Panobinostat has been FDA-approved in combination with bortezomib and dexamethasone for RRMM. Preclinical data show that selective HDAC6 inhibition (e.g., ricolinostat) may enhance the efficacy of BET inhibitors like JQ1 by restoring MYC suppression and promoting apoptosis [229]. Importantly, dual HDAC/BET inhibitors are in development and show greater efficacy than either agent alone, with reduced off-target toxicity in preclinical models [230]. EZH2, a methyltransferase in the PRC2 complex, mediates trimethylation of histone H3K27 and represses gene expression. EZH2 is frequently overexpressed in MM and associated with poor prognosis and cell cycle dysregulation [86]. Although EZH2 inhibitors (e.g., GSK126, EPZ-6438) have limited single-agent activity in MM, their combination with HDAC inhibitors (like panobinostat) results in synergistic cytotoxicity and extensive transcriptional reprogramming that promotes apoptosis [87]. Furthermore, triple combinations targeting BMI-1 (PRC1), EZH2 (PRC2), and BET proteins have shown additive or synergistic anti-myeloma effects, enhancing chromatin accessibility and inducing durable apoptosis in vitro [90].

7.3. Vaccine Strategies and Tumor Neoantigens

Cancer vaccines targeting tumor-specific neoantigens represent an emerging and highly personalized immunotherapeutic strategy in MM (Table 17), particularly as a complement to existing T-cell-based therapies. Neoantigens—arising from somatic mutations in tumor cells—are not present in normal tissue and therefore evade central tolerance, making them ideal immunologic targets for vaccine development. Preclinical and early clinical studies in MM have confirmed that neoantigen-based vaccines can elicit both CD4+ and CD8+ T-cell responses, with protective and therapeutic potential. In a mouse model of MM (MOPC315), a vaccine composed of an idiotype (Id) neoantigen fused with a heat shock binding motif and poly (I:C) adjuvant induced strong CD4+ T-helper responses, which in turn activated CD8+ cytotoxic T cells against non-vaccine tumor-associated antigens, resulting in tumor control and immune memory [231]. Neoantigen vaccines are now entering clinical trials. In the PGV-001 phase I trial, a multipeptide personalized neoantigen vaccine was administered to patients (including those with MM) in the post-remission setting. Each patient’s tumor underwent genomic sequencing to identify 8–10 predicted MHC class I-binding neoantigens. The vaccine induced neoantigen-specific CD4 and CD8 T-cell responses, was well tolerated, and confirmed the feasibility of individualized peptide vaccine synthesis in MM [232,233]. Furthermore, neoantigen-specific T-cell responses have now been validated directly in MM patients. A study analyzing tumor sequencing data from 184 MM patients identified shared immunogenic mutations in driver genes like KRAS, NRAS, and IRF4 and demonstrated that neoantigen-specific CD8+ T cells correlated with tumor regression in relapsed patients [234]. These results suggest that even in cancers with intermediate mutational burden, like MM, neoantigen-based vaccines are feasible and may enhance immune surveillance.
Next, a phase I clinical trial evaluated the safety and immunogenicity of therapeutic peptide vaccines targeting Bcl-2 family proteins (Bcl-2, Bcl-XL, Mcl-1) in patients with relapsed MM. Seven patients received vaccinations alongside bortezomib, with four completing the full eight-dose schedule. The vaccine was well tolerated, with no unexpected toxicities beyond those associated with bortezomib. Among the six patients who received more than two doses, immune responses to the peptides were observed in all, and three showed increased immune reactivity after vaccination. These results suggest that Bcl-2-targeted peptide vaccination is feasible, safe, and immunologically active in relapsed MM, supporting further investigation in larger trials [235]. Another study evaluated the immunologic potential of a multipeptide vaccine cocktail composed of four HLA-A2–restricted peptides derived from MM-associated antigens—XBP1 (US184–192 and SP367–375), CD138 (260–268), and CS1 (239–247). Cytotoxic T lymphocytes (CTLs) generated from HLA-A2+ donors using this peptide cocktail demonstrated robust, antigen-specific immune responses, including increased populations of effector memory (CCR7CD45RO+) and activated (CD69+) CD8+ T cells. The multipronged CTLs also showed proliferation, IFN-γ production, and cytotoxic activity against HLA-A2+ MM cell lines and primary MM patient cells, while sparing irrelevant targets. Importantly, each peptide within the cocktail contributed to antigen-specific activity without compromising overall immune function. These findings support the feasibility of using multiantigen peptide vaccines to generate broad and potent MM-specific T-cell responses [236].
Innovative formulations are under investigation to improve immune activation. For example, synthetic peptide vaccines incorporating heat shock motifs or delivered via viral or nanoparticle platforms are being tested to improve antigen presentation and elicit broader, durable T-cell responses [237]. Combinations with checkpoint inhibitors or IMiDs are also being explored to overcome T-cell exhaustion and enhance vaccine potency [238]. One challenge remains the heterogeneity of neoantigens between patients and within tumor subclones, which necessitates highly personalized vaccine design. Advances in sequencing and neoantigen prediction pipelines (e.g., OpenVax) are enabling rapid identification and ranking of immunogenic targets for vaccine synthesis.

7.4. Microenvironment Modulation

The bone marrow microenvironment in MM plays a pivotal role in promoting tumor progression, immune evasion, and therapy resistance. Key players in this immunosuppressive niche include myeloid-derived suppressor cells (MDSCs), Tregs, and dysregulated cytokine networks. These elements foster a tumor-promoting immune landscape, thereby undermining the efficacy of immunotherapies and contributing to relapse. MDSCs are significantly expanded in both the peripheral blood and bone marrow of MM patients, particularly in those with active disease. They not only enhance MM cell survival and proliferation but also suppress T-cell responses via arginase-1, reactive oxygen species (ROS), and inhibitory cytokines like IL-6 and IL-10 [239]. Furthermore, MDSCs facilitate the induction of Tregs and impair dendritic cell function, leading to broader immunosuppression and reduced antitumor surveillance [240]. Therapeutic strategies targeting MDSCs are under active investigation. Notably, the multi-kinase inhibitor sunitinib has shown efficacy in suppressing both monocytic and neutrophilic MDSC subsets, restoring T-cell proliferation and enhancing the cytotoxic effects of lenalidomide in co-culture models [241]. Additionally, HDAC inhibitors and phosphodiesterase-5 (PDE5) inhibitors are being evaluated for their capacity to deplete or reprogram MDSCs [242]. Tregs also accumulate in MM and correlate with poor prognosis. These FoxP3+ cells suppress effector T-cell activation and facilitate immune escape. Targeting Tregs—either through IL-2-diphtheria toxin fusion proteins (e.g., E7777) or checkpoint blockade combinations—has shown promise in reducing immune suppression and enhancing tumor-specific responses [243]. Moreover, cytokine dysregulation, especially overproduction of IL-6, IL-10, TGF-β, and prostaglandin E2 (PGE2), reinforces the immunosuppressive loop within the bone marrow microenvironment. IL-6 promotes plasma cell survival and drug resistance, while IL-10 and TGF-β inhibit cytotoxic T cells and dendritic cell function. Targeting these cytokines or their receptors (e.g., anti-IL-6 monoclonal antibodies or EP4 antagonists) is being explored to restore immune balance and improve responses to immunotherapy [243,244].

7.5. Precision Medicine and Biomarker-Driven Trials

Precision medicine in MM aims to tailor therapies based on individual molecular and clinical profiles, moving beyond the traditional “one-size-fits-all” paradigm. With the rise of NGS and multi-omics profiling, clinical trial designs have evolved to accommodate this complexity. Among these, basket and umbrella trials (Figure 4)—core master protocol strategies—enable more adaptive and biomarker-driven evaluation of novel therapies. Basket trials enroll patients with different tumor types but a common molecular alteration, allowing researchers to assess whether a therapy targeting that alteration is effective regardless of the cancer’s tissue origin. In contrast, umbrella trials focus on a single tumor type (e.g., MM) but test multiple therapies matched to different molecular subtypes within that disease. These trial designs allow efficient stratification, faster drug development, and more personalized therapy allocation [245,246]. In MM, where inter- and intra-patient heterogeneity is high, umbrella trials could enable simultaneous testing of targeted agents against genomic aberrations like t(11;14), del(17p), or RAS mutations in defined subgroups. However, the success of these designs hinges on a clear understanding of biomarker classification. Prognostic biomarkers provide information about a patient’s likely disease outcome independent of treatment. For example, del(17p) and gain(1q21) are negative prognostic markers in MM, predicting early relapse and shorter OS. In contrast, predictive biomarkers indicate response likelihood to a specific therapy, enabling patient selection for targeted treatments. BCL2 expression in t(11;14) myeloma, which predicts response to venetoclax, is a classic example of a predictive biomarker [247].
Many biomarker-driven trials now incorporate adaptive enrichment strategies, in which ongoing biomarker assessments refine trial arms and eligibility as data accumulates. These dynamic designs are critical in MM, where clonal evolution may alter mutational profiles over time. Additionally, platform trials—a hybrid of basket and umbrella designs—allow continuous addition or closure of arms based on interim efficacy data, making them well-suited for rapidly evolving therapeutic landscapes like MM [248]. Despite these advances, challenges persist. The combinatorial complexity of MM mutations, subclonal heterogeneity, and overlap of predictive and prognostic functions complicate trial interpretation and therapy matching. Moreover, robust biomarker validation is often lacking in early-phase MM studies, limiting their generalizability [249].

8. Special Considerations

8.1. Management in Elderly and Frail Patients

MM is predominantly a disease of older adults, with a median age at diagnosis of approximately 70 years. As the aging population grows, so does the clinical challenge of managing MM in elderly and frail patients, who often present with multimorbidity, reduced functional reserve, cognitive decline, and polypharmacy. These factors contribute to increased treatment-related toxicity, early therapy discontinuation, and worse survival outcomes compared to younger or fitter patients [250,251]. Historically, treatment selection in elderly patients was based on chronological age and performance status. However, age alone is insufficient to predict tolerability. Instead, frailty assessments using tools such as IMWG frailty index and the Revised Myeloma Comorbidity Index (R-MCI) are now recommended. These tools stratify patients into fit, intermediate-fit, or frail, guiding treatment intensity accordingly [252]. Frailty-adapted therapy aims to balance efficacy with toxicity, prioritizing symptom control, quality of life, and functional independence. Fit elderly patients may receive full-dose triplet regimens like VRd, while frail individuals often benefit more from two-drug combinations or attenuated regimens, such as dose-reduced lenalidomide and dexamethasone (Rd-lite) [253]. Newer oral triplet regimens like ixazomib-Rd have also shown promise in frail patients, offering lower toxicity and improved quality of life [254].
Evidence from real-world studies and prospective trials supports the prognostic value of frailty assessments, not only at baseline but also as a dynamic metric during therapy. A large meta-analysis demonstrated that frailty independently predicts mortality and adverse events in MM regardless of treatment choice [255]. Moreover, routine geriatric reassessments can help adjust therapy over time and identify patients at risk of treatment discontinuation due to evolving impairments [256]. To facilitate implementation in daily practice, electronic and rapid-screening tools are being integrated into clinical workflows. These digital geriatric assessments require minimal time and have been shown to influence treatment decisions in over 30% of cases, with clinicians escalating or de-escalating therapy based on frailty status [257]. As such, comprehensive geriatric assessment (CGA) is essential in guiding individualized, frailty-adapted therapy for elderly MM patients. It improves therapeutic decision-making, minimizes toxicity, and aligns treatment with patient-centered goals, marking a critical step toward precision oncology in aging populations.

8.2. Racial and Ethnic Disparities

MM presents significant racial and ethnic disparities in both incidence and outcomes. African Americans (AAs) have a two- to threefold higher incidence of MM compared to White Americans, and they tend to be diagnosed at younger ages [258]. Despite this increased risk, some studies have found better or equivalent OS among AAs when provided with equal access to care, suggesting that non-biological factors, such as access to treatment and healthcare delivery systems, may be key contributors to outcome disparities [259]. Recent genomic studies have uncovered differences in disease biology that may partially explain disparities. AAs exhibit a higher frequency of the t(11;14) translocation and a lower prevalence of high-risk cytogenetic abnormalities such as del(17p), which could contribute to relatively favorable biology in this group [260]. Furthermore, data from the National Institute of Health (NIH) “All of Us” cohort demonstrated a synergistic interaction between African genetic ancestry and hypertension as a strong risk factor for MM, highlighting the role of gene–environment interplay in racial disparities [261].
Yet, the social determinants of health (SDOH) remain the dominant modifiable contributors to disparities. Socioeconomic status, insurance coverage, education level, and geographic access to specialized care profoundly affect outcomes. For example, Black and Hispanic patients are less likely to receive ASCT and novel agents like lenalidomide or monoclonal antibodies, even when clinically indicated [262,263]. These treatment gaps are further exacerbated by underrepresentation of minorities in clinical trials, limiting generalizability and access to cutting-edge therapies [264]. Moreover, Hispanic and Native American patients face additional barriers and often present with more advanced disease. A study using SEER data revealed that Hispanic patients had the worst OS, potentially due to both healthcare access issues and differences in disease biology [265]. Among Hispanics, younger age at diagnosis and higher rates of extramedullary disease were reported, but cytogenetic profiles did not clearly explain survival gaps, reinforcing the need to further explore healthcare inequities [266]. In conclusion, racial and ethnic disparities in MM reflect a complex interplay between genetic predisposition and socioeconomic disadvantage. Addressing these disparities will require not only a deeper understanding of disease biology across ancestral backgrounds but also systemic reform to ensure equitable access to diagnostics, therapies, and clinical trials.

8.3. Relapsed and Refractory MM

RRMM presents a significant therapeutic challenge, as nearly all patients eventually relapse despite initial responses to frontline therapy. The goal in RRMM is to maximize disease control while preserving quality of life through the rational use of sequential, non-cross-resistant therapies and incorporation of emerging agents. Therapy selection is increasingly individualized based on prior treatments, depth and duration of previous response, patient comorbidities, and relapse aggressiveness. For patients with early relapse (e.g., <12 months after autologous stem cell transplant or triple-class exposure), intensive triplet or quadruplet rescue regimens are preferred, often including next-generation PIs (carfilzomib), immunomodulatory drugs (pomalidomide), or monoclonal antibodies (daratumumab, isatuximab) [267]. Patients with indolent relapse or frailty may benefit from reusing previously effective regimens, particularly if the treatment-free interval exceeded 18 months [268]. Sequential therapy, rather than upfront combination of all active agents, remains a viable strategy in certain scenarios, particularly for chronic management of RRMM. This approach reserves newer agents for future lines, potentially reducing cumulative toxicity and extending overall treatment durability. However, combination therapy in early relapse may lead to deeper responses and prolonged survival—highlighting the need to balance treatment intensity with patient-specific factors [269]. Clinical trials are pivotal in the RRMM setting, especially for patients who are refractory to PIs, IMiDs, and anti-CD38 antibodies—so-called triple-class refractory disease. Participation in trials offers access to novel mechanisms of action, such as bispecific antibodies, CAR-T cells, ADCs (e.g., belantamab mafodotin), and CELMoDs, many of which have demonstrated promising activity in early-phase studies [270]. For example, patients enrolled in early trials combining lenalidomide and bortezomib achieved durable responses with manageable toxicity even after multiple relapses [271]. Importantly, clinical trial participation improves survival outcomes compared to standard therapies alone and should be strongly encouraged, especially when optimal treatment sequences remain undefined [272]. With an increasing number of trials incorporating biomarker stratification and real-time molecular monitoring, future RRMM therapy will likely become even more personalized and adaptive.

9. Conclusions

MM represents a biologically complex and clinically heterogeneous malignancy characterized by progressive genetic and epigenetic alterations, immune escape, and microenvironmental crosstalk. Over the past two decades, advances in molecular diagnostics, risk stratification, and the development of targeted and immune-based therapies have significantly improved patient outcomes. However, MM remains incurable for most patients, with inevitable relapse and the emergence of drug resistance posing ongoing challenges. Looking ahead, future strategies must focus on integrating genomic and immune profiling into real-time clinical decision-making to enable personalized therapy selection. Emphasis should be placed on developing biomarker-driven clinical trials, especially adaptive basket and umbrella designs, to match emerging agents with molecularly defined subgroups. The microenvironment and mechanisms of immune suppression, including the roles of MDSCs, Tregs, and cytokine networks, remain promising therapeutic targets. Moreover, optimizing the use of CAR-T cells, bispecific T-cell engagers, and vaccine-based approaches in combination regimens or earlier disease stages holds potential for deeper and more durable responses. As treatment becomes increasingly complex, frailty assessments, equity in access, and inclusive trial enrollment will be critical to ensure that novel advances benefit all patient populations. Ultimately, a multidisciplinary, precision-guided approach is essential to transform MM into a manageable—and eventually curable—disease.

Author Contributions

Conceptualization, V.R.F., P.Z.R. and H.S.; methodology, H.S.; validation, H.S.; formal analysis, H.S.; investigation, P.Z.R. and H.S.; resources, H.S.; data curation, V.R.F., P.Z.R. and H.S.; writing—original draft preparation, V.R.F., P.Z.R., H.S., A.W., B.A.A., B.R., E.E., L.P. and G.J.A.; writing—review and editing, V.R.F., P.Z.R., H.S., A.W., B.A.A., B.R., E.E., L.P. and G.J.A.; visualization, H.S.; project administration, H.S.; funding acquisition, H.S.; supervision, P.Z.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

No new data were created.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Mateos, M.-V.; Landgren, O. MGUS and Smoldering Multiple Myeloma: Diagnosis and Epidemiology. In Plasma Cell Dyscrasias; Roccaro, A.M., Ghobrial, I.M., Eds.; Springer International Publishing: Cham, Switzerland, 2016; pp. 3–12. ISBN 978-3-319-40320-5. [Google Scholar]
  2. Sutanto, H.; Sandra, D.Y.; Safira, A.; Adytia, G.J.; Waitupu, A.; Romadhon, P.Z. Genetic, Epigenetic, and Molecular Determinants of Multiple Myeloma and Precursor Plasma Cell Disorders: A Pathophysiological Overview. Med. Oncol. 2025, 42, 234. [Google Scholar] [CrossRef] [PubMed]
  3. Van De Donk, N.W.C.J.; Mutis, T.; Poddighe, P.J.; Lokhorst, H.M.; Zweegman, S. Diagnosis, Risk Stratification and Management of Monoclonal Gammopathy of Undetermined Significance and Smoldering Multiple Myeloma. Int. J. Lab. Hematol. 2016, 38, 110–122. [Google Scholar] [CrossRef] [PubMed]
  4. González-Calle, V.; Mateos, M.V. Monoclonal Gammopathies of Unknown Significance and Smoldering Myeloma: Assessment and Management of the Elderly Patients. Eur. J. Intern. Med. 2018, 58, 57–63. [Google Scholar] [CrossRef] [PubMed]
  5. Ho, M.; Patel, A.; Goh, C.Y.; Moscvin, M.; Zhang, L.; Bianchi, G. Changing Paradigms in Diagnosis and Treatment of Monoclonal Gammopathy of Undetermined Significance (MGUS) and Smoldering Multiple Myeloma (SMM). Leukemia 2020, 34, 3111–3125. [Google Scholar] [CrossRef]
  6. Korde, N.; Kristinsson, S.Y.; Landgren, O. Monoclonal Gammopathy of Undetermined Significance (MGUS) and Smoldering Multiple Myeloma (SMM): Novel Biological Insights and Development of Early Treatment Strategies. Blood 2011, 117, 5573–5581. [Google Scholar] [CrossRef]
  7. Blum, A.; Bazou, D.; O’Gorman, P. Smoldering Multiple Myeloma: Prevalence and Current Evidence Guiding Treatment Decisions. BLCTT 2018, 8, 21–31. [Google Scholar] [CrossRef]
  8. 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]
  9. Furukawa, Y.; Kikuchi, J. Molecular Basis of Clonal Evolution in Multiple Myeloma. Int. J. Hematol. 2020, 111, 496–511. [Google Scholar] [CrossRef]
  10. Salomon-Perzyński, A.; Jamroziak, K.; Głodkowska-Mrówka, E. Clonal Evolution of Multiple Myeloma—Clinical and Diagnostic Implications. Diagnostics 2021, 11, 1534. [Google Scholar] [CrossRef]
  11. Qin, X.; An, G.; Hu, L.; Qin, Y.; Xu, Y.; Feng, X.; Zang, M.; Deng, S.; Sui, W.; Yi, S.; et al. Clonal Evolution Paths Significantly Impact the Outcome of Myeloma. Blood 2015, 126, 1784. [Google Scholar] [CrossRef]
  12. Yan, Y.; Qin, X.; Liu, J.; Fan, H.; Yu, Z.; Liu, W.; Hao, M.; Qiu, L.; An, G. OAB-011: Clonal Phylogeny and Evolution of Critical Cytogenetic Aberrations in Multiple Myeloma at Single Cell Level. Clin. Lymphoma Myeloma Leuk. 2021, 21, S7–S8. [Google Scholar] [CrossRef]
  13. Farswan, A.; Jena, L.; Kaur, G.; Gupta, A.; Gupta, R.; Rani, L.; Sharma, A.; Kumar, L. Branching Clonal Evolution Patterns Predominate Mutational Landscape in Multiple Myeloma. Am. J. Cancer Res. 2021, 11, 5659–5679. [Google Scholar] [PubMed]
  14. Melchor, L.; Brioli, A.; Wardell, C.P.; Murison, A.; Potter, N.E.; Kaiser, M.F.; Fryer, R.A.; Johnson, D.C.; Begum, D.B.; Hulkki Wilson, S.; 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]
  15. Suzuki, K.; Nishiwaki, K.; Yano, S. Treatment Strategies Considering Micro-Environment and Clonal Evolution in Multiple Myeloma. Cancers 2021, 13, 215. [Google Scholar] [CrossRef] [PubMed]
  16. 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 2019, 33, 457–468. [Google Scholar] [CrossRef]
  17. Van Nieuwenhuijzen, N.; Spaan, I.; Raymakers, R.; Peperzak, V. From MGUS to Multiple Myeloma, a Paradigm for Clonal Evolution of Premalignant Cells. Cancer Res. 2018, 78, 2449–2456. [Google Scholar] [CrossRef]
  18. Bianchi, G.; Ghobrial, I. Biological and Clinical Implications of Clonal Heterogeneity and Clonal Evolution in Multiple Myeloma. CCTR 2014, 10, 70–79. [Google Scholar] [CrossRef]
  19. Corsale, A.M.; Di Simone, M.; Shekarkar Azgomi, M.; Speciale, M.; Gigliotta, E.; Garofano, F.; Aquilina, C.; Vullo, C.; Caccamo, N.; Dieli, F.; et al. Immune Cells, Cytokines, and Gut Microbiota Landscape Along Monoclonal Gammopathy of Undetermined Significance (MGUS) to Multiple Myeloma (MM) Evolution. Blood 2023, 142, 3292. [Google Scholar] [CrossRef]
  20. Manier, S.; Sacco, A.; Leleu, X.; Ghobrial, I.M.; Roccaro, A.M. Bone Marrow Microenvironment in Multiple Myeloma Progression. J. Biomed. Biotechnol. 2012, 2012, 157496. [Google Scholar] [CrossRef]
  21. Cenzano, I.; Cócera, M.; Larrayoz, M.; Campos-Dopazo, L.; Sanz, S.; Bantan, A.; Vilas-Zornoza, A.; San-Martin, P.; Aguirre-Ruiz, P.; Alignani, D.; et al. Stromal and Endothelial Transcriptional Changes during Progression from MGUS to Myeloma and after Treatment Response. bioRxiv 2024. [Google Scholar] [CrossRef]
  22. Capp, J.-P.; Bataille, R. The Ins and Outs of Endosteal Niche Disruption in the Bone Marrow: Relevance for Myeloma Oncogenesis. Biology 2023, 12, 990. [Google Scholar] [CrossRef]
  23. Lopes, R.; Caetano, J.; Ferreira, B.; Barahona, F.; Carneiro, E.A.; João, C. The Immune Microenvironment in Multiple Myeloma: Friend or Foe? Cancers 2021, 13, 625. [Google Scholar] [CrossRef]
  24. Rees, E.; Sodi, I.; Foster, K.; Ainley, L.; Lyon, E.; Galas-Filipowicz, D.; Rahman, J.; Allen, R.; Kimber, J.; Prabu, A.; et al. Single-Cell Profiling of Human Bone Marrow Reveals Multiple Myeloma Progression Is Accompanied by an Increase in CD56brightBone Marrow Resident NK Cells. bioRxiv 2025. [Google Scholar] [CrossRef]
  25. Janker, L.; Mayer, R.L.; Bileck, A.; Kreutz, D.; Mader, J.C.; Utpatel, K.; Heudobler, D.; Agis, H.; Gerner, C.; Slany, A. Metabolic, Anti-Apoptotic and Immune Evasion Strategies of Primary Human Myeloma Cells Indicate Adaptations to Hypoxia*. Mol. Cell. Proteom. 2019, 18, 936a–953. [Google Scholar] [CrossRef]
  26. Alagpulinsa, D.A.; Szalat, R.E.; Poznansky, M.C.; Shmookler Reis, R.J. Genomic Instability in Multiple Myeloma. Trends Cancer 2020, 6, 858–873. [Google Scholar] [CrossRef]
  27. Mahtouk, K.; Hose, D.; Rème, T.; De Vos, J.; Jourdan, M.; Moreaux, J.; Fiol, G.; Raab, M.; Jourdan, E.; Grau, V.; et al. Expression of EGF-Family Receptors and Amphiregulin in Multiple Myeloma. Amphiregulin Is a Growth Factor for Myeloma Cells. Oncogene 2005, 24, 3512–3524. [Google Scholar] [CrossRef] [PubMed]
  28. Saltarella, I.; Apollonio, B.; Lamanuzzi, A.; Desantis, V.; Mariggiò, M.A.; Desaphy, J.-F.; Vacca, A.; Frassanito, M.A. The Landscape of lncRNAs in Multiple Myeloma: Implications in the “Hallmarks of Cancer”, Clinical Perspectives and Therapeutic Opportunities. Cancers 2022, 14, 1963. [Google Scholar] [CrossRef] [PubMed]
  29. Bolzoni, M.; Chiu, M.; Accardi, F.; Vescovini, R.; Airoldi, I.; Storti, P.; Todoerti, K.; Agnelli, L.; Missale, G.; Andreoli, R.; et al. Dependence on Glutamine Uptake and Glutamine Addiction Characterize Myeloma Cells: A New Attractive Target. Blood 2016, 128, 667–679. [Google Scholar] [CrossRef] [PubMed]
  30. Bajpai, R.; Matulis, S.M.; Wei, C.; Nooka, A.K.; Von Hollen, H.E.; Lonial, S.; Boise, L.H.; Shanmugam, M. Targeting Glutamine Metabolism in Multiple Myeloma Enhances BIM Binding to BCL-2 Eliciting Synthetic Lethality to Venetoclax. Oncogene 2016, 35, 3955–3964. [Google Scholar] [CrossRef]
  31. Soncini, D.; Minetto, P.; Martinuzzi, C.; Becherini, P.; Fenu, V.; Guolo, F.; Todoerti, K.; Calice, G.; Contini, P.; Miglino, M.; et al. Amino Acid Depletion Triggered by l-Asparaginase Sensitizes MM Cells to Carfilzomib by Inducing Mitochondria ROS-Mediated Cell Death. Blood Adv. 2020, 4, 4312–4326. [Google Scholar] [CrossRef]
  32. Effenberger, M.; Bommert, K.S.; Kunz, V.; Kruk, J.; Leich, E.; Rudelius, M.; Bargou, R.; Bommert, K. Glutaminase Inhibition in Multiple Myeloma Induces Apoptosis via MYC Degradation. Oncotarget 2017, 8, 85858–85867. [Google Scholar] [CrossRef] [PubMed]
  33. Nadav, L.; Katz, B.-Z.; Baron, S.; Cohen, N.; Naparstek, E.; Geiger, B. The Generation and Regulation of Functional Diversity of Malignant Plasma Cells. Cancer Res. 2006, 66, 8608–8616. [Google Scholar] [CrossRef] [PubMed]
  34. Loiseau, H.A.; Li, C.; Magrangeas, F.; Guerin, C.; Gouraud, W.; Harousseau, J.-L.; Attal, M.; Marit, G.; Mathiot, C.; Facon, T.; et al. In Myeloma, the Prognostic Impact of Hyperdiploidy Is Mainly Related to the Gain of Chromosome 5. Blood 2008, 112, 632. [Google Scholar] [CrossRef]
  35. Kaiser, M.F.; Boyle, E.M.; Walker, B.A.; Begum, D.B.; Proszek, P.; Johnson, D.C.; Pawlyn, C.; Jones, J.R.; Savola, S.; Owen, R.G.; et al. Molecular Subgroups of Hyperdiploidy and Their Prognostic Relevance—An Analysis of 1036 Myeloma Trial Patients. Blood 2015, 126, 2983. [Google Scholar] [CrossRef]
  36. Chng, W.J.; Kumar, S.; VanWier, S.; Ahmann, G.; Price-Troska, T.; Henderson, K.; Chung, T.-H.; Kim, S.; Mulligan, G.; Bryant, B.; et al. Molecular Dissection of Hyperdiploid Multiple Myeloma by Gene Expression Profiling. Cancer Res. 2007, 67, 2982–2989. [Google Scholar] [CrossRef]
  37. AlHashmi, H.; Al-Dayel, A.; Soliman, D.; Al-Sayegh, M.; Abduljalil, O.; Anezi, K.A.; Matrok, A.A.; Kaloyannidis, P.; Saber, A.A.; Kamel, M.M.; et al. Hyperdiploidy Is a Positive Prognostic Factor for Progression-Free Survival in Multiple Myeloma with High and Intermediate Risk Cytogenetics. Health Sci. J. 2018, 12, 590. [Google Scholar] [CrossRef]
  38. Pawlyn, C.; Melchor, L.; Murison, A.; Wardell, C.P.; Brioli, A.; Boyle, E.M.; Kaiser, M.F.; Walker, B.A.; Begum, D.B.; Dahir, N.B.; et al. Coexistent Hyperdiploidy Does Not Abrogate Poor Prognosis in Myeloma with Adverse Cytogenetics and May Precede IGH Translocations. Blood 2015, 125, 831–840. [Google Scholar] [CrossRef]
  39. Agnelli, L.; Fabris, S.; Bicciato, S.; Basso, D.; Baldini, L.; Morabito, F.; Verdelli, D.; Todoerti, K.; Lambertenghi-Deliliers, G.; Lombardi, L.; et al. Upregulation of Translational Machinery and Distinct Genetic Subgroups Characterise Hyperdiploidy in Multiple Myeloma. Br. J. Haematol. 2007, 136, 565–573. [Google Scholar] [CrossRef]
  40. Weinhold, N.; DeVos, J.; Hose, D.; Rossi, J.-F.; Axel, B.; Mahtouk, K.; Raab, M.S.; Rème, T.X.; Jauch, A.; Moreaux, J.; et al. Ribosomal Proteins Are Overexpressed in Hyperdiploid Multiple Myeloma. Blood 2007, 110, 2495. [Google Scholar] [CrossRef]
  41. Neri, A.; Todoerti, K.; Agnelli, L.; Fabris, S.; Bicciato, S.; Mosca, L.; Nobili, L.; Verdelli, D.; Ronchetti, D.; Intini, D.; et al. Identification of Specific Transcriptional Patterns Associated with Hyperdiploidy in Multiple Myeloma. Blood 2006, 108, 3412. [Google Scholar] [CrossRef]
  42. Rio-Machin, A.; Ferreira, B.I.; Henry, T.; Gómez-López, G.; Agirre, X.; Alvarez, S.; Rodriguez-Perales, S.; Prosper, F.; Calasanz, M.J.; Martínez, J.; et al. Downregulation of Specific miRNAs in Hyperdiploid Multiple Myeloma Mimics the Oncogenic Effect of IgH Translocations Occurring in the Non-Hyperdiploid Subtype. Leukemia 2013, 27, 925–931. [Google Scholar] [CrossRef]
  43. Samur, M.K.; Aktas Samur, A.; Shah, P.; Park, J.S.; Fulciniti, M.; Shammas, M.; Corre, J.; Anderson, K.C.; Parmigiani, G.; Avet-Loiseau, H.; et al. Development of Hyperdiploidy Starts at an Early Age and Takes a Decade to Complete. Blood 2025, 145, 520–525. [Google Scholar] [CrossRef]
  44. Miura, D.; Narita, K.; Kuzume, A.; Tabata, R.; Terao, T.; Tsushima, T.; Kobayashi, H.; Abe, Y.; Kitadate, A.; Takeuchi, M.; et al. Clinical and Prognostic Impact of (11;14)(Q13;Q32) Translocation on Patients with Multiple Myeloma. Blood 2019, 134, 5507. [Google Scholar] [CrossRef]
  45. Chakraborty, R.; Lentzsch, S. Prognostic Impact of t(11;14) in Multiple Myeloma: Black and White or Shades of Gray? Cancer 2021, 127, 31–34. [Google Scholar] [CrossRef] [PubMed]
  46. Chang, H.; Sloan, S.; Li, D.; Zhuang, L.; Yi, Q.; Chen, C.I.; Reece, D.; Chun, K.; Keith Stewart, A. The t(4;14) Is Associated with Poor Prognosis in Myeloma Patients Undergoing Autologous Stem Cell Transplant. Br. J. Haematol. 2004, 125, 64–68. [Google Scholar] [CrossRef] [PubMed]
  47. Sato, S.; Kamata, W.; Okada, S.; Tamai, Y. Clinical and Prognostic Significance of t(4;14) Translocation in Multiple Myeloma in the Era of Novel Agents. Int. J. Hematol. 2021, 113, 207–213. [Google Scholar] [CrossRef]
  48. Walker, B.A.; Wardell, C.P.; Murison, A.; Boyle, E.M.; Melchor, L.; Pawlyn, C.; Kaiser, M.F.; Begum, D.; Dahir, N.; Proszek, P.; et al. Apobec Family Mutational Signatures Are Associated with Poor Prognosis Translocations in Multiple Myeloma. Blood 2014, 124, 723. [Google Scholar] [CrossRef]
  49. Marzocchi, G.; Ameli, G.; Pezzi, A.; Zamagni, E.; Cavallo, F.; Petrucci, M.T.; Waage, A.; Di Raimondo, F.; Patriarca, F.; Rambaldi, A.; et al. Rare Igh Translocations in Newly Diagnosed Multiple Myeloma (MM) Patients: Cytogenetic Characterization and Relevance on Prognosis. Blood 2014, 124, 2042. [Google Scholar] [CrossRef]
  50. Mao, X.; Zhuang, J.; Zhao, D.; Li, X.; Du, X.; Hao, M.; Xu, Y.; Yan, Y.; Liu, J.; Fan, H.; et al. IgH Translocation with Undefined Partners Is Associated with Superior Outcome in Multiple Myeloma Patients. Eur. J. Haematol. 2020, 105, 326–334. [Google Scholar] [CrossRef]
  51. Fonseca, R.; Oken, M.; Harrington, D.; Bailey, R.; Van Wier, S.; Henderson, K.; Kay, N.; Van Ness, B.; Greipp, P.; Dewald, G. Deletions of Chromosome 13 in Multiple Myeloma Identified by Interphase FISH Usually Denote Large Deletions of the q Arm or Monosomy. Leukemia 2001, 15, 981–986. [Google Scholar] [CrossRef]
  52. Kaufmann, H.; Krömer, E.; Nösslinger, T.; Weltermann, A.; Ackermann, J.; Reisner, R.; Bernhart, M.; Drach, J. Both Chromosome 13 Abnormalities by Metaphase Cytogenetics and Deletion of 13q by Interphase FISH Only Are Prognostically Relevant in Multiple Myeloma. Eur. J. Haematol. 2003, 71, 179–183. [Google Scholar] [CrossRef]
  53. Fassas, A.B.-T.; Spencer, T.; Sawyer, J.; Zangari, M.; Lee, C.; Anaissie, E.; Muwalla, F.; Morris, C.; Barlogie, B.; Tricot, G. Both Hypodiploidy and Deletion of Chromosome 13 Independently Confer Poor Prognosis in Multiple Myeloma. Br. J. Haematol. 2002, 118, 1041–1047. [Google Scholar] [CrossRef] [PubMed]
  54. Li, Q.; An, G.; Li, C.; Xu, Y.; Deng, S.; Liu, X.; Qi, J.; Wang, Y.; Zou, D.; Zhao, Y.; et al. The significances of 13q14 deletion for development and prognosis of multiple myeloma. Zhonghua Xue Ye Xue Za Zhi 2011, 32, 217–220. [Google Scholar] [PubMed]
  55. Avet-Loiseau, H.; Leleu, X.; Roussel, M.; Mathiot, C.; Caillot, D.; Hulin, C.; Marit, G.; Facon, T.; Attal, M.; Harousseau, J.-L.; et al. Deletion of the 17p Chromosomal Region Is Associated with a Very Poor Outcome in Multiple Myeloma Independently of the Type of Treatment. Blood 2009, 114, 1817. [Google Scholar] [CrossRef]
  56. Lode, L.; Eveillard, M.; Trichet, V.; Soussi, T.; Wuilleme, S.; Richebourg, S.; Magrangeas, F.; Ifrah, N.; Campion, L.; Traulle, C.; et al. Mutations in TP53 Are Exclusively Associated with Del(17p) in Multiple Myeloma. Haematologica 2010, 95, 1973–1976. [Google Scholar] [CrossRef]
  57. Han, E.; Kwon, A.; Kim, Y.; Han, K.; Park, S.; Min, C.; Kim, M. Prevalence of Ras/Mapk Pathway Mutation (KRAS, NRAS, and BRAF) in Plasma Cell Myeloma at a Single Institute in Korea. Cytotherapy 2019, 21, S25–S26. [Google Scholar] [CrossRef]
  58. Janku, F.; Li, Q.; Huang, H.J.; Wang, Y.; Cao, Z.; Karlin-Neumann, G.; Liu, Z. Abstract 573: BRAF, KRAS, and NRAS Mutations in Archival Tumor Samples and Samples of Cell-Free DNA from Serum and Bone Marrow Aspirates from Patients with Multiple Myeloma. Cancer Res. 2018, 78, 573. [Google Scholar] [CrossRef]
  59. Mulligan, G.; Lichter, D.I.; Di Bacco, A.; Blakemore, S.J.; Berger, A.; Koenig, E.; Bernard, H.; Trepicchio, W.; Li, B.; Neuwirth, R.; et al. Mutation of NRAS but Not KRAS Significantly Reduces Myeloma Sensitivity to Single-Agent Bortezomib Therapy. Blood 2014, 123, 632–639. [Google Scholar] [CrossRef]
  60. Kim, Y.; Park, S.-S.; Min, C.-K.; Lee, G.D.; Son, J.; Jo, S.J.; Han, E.; Han, K.; Kim, M. KRAS, NRAS, and BRAF Mutations in Plasma Cell Myeloma at a Single Korean Institute. Blood Res. 2020, 55, 159–168. [Google Scholar] [CrossRef]
  61. Shirazi, F.; Jones, R.J.; Singh, R.K.; Zou, J.; Kuiatse, I.; Berkova, Z.; Wang, H.; Lee, H.C.; Hong, S.; Dick, L.; et al. Activating KRAS, NRAS, and BRAF Mutants Enhance Proteasome Capacity and Reduce Endoplasmic Reticulum Stress in Multiple Myeloma. Proc. Natl. Acad. Sci. USA 2020, 117, 20004–20014. [Google Scholar] [CrossRef]
  62. Lionetti, M.; Barbieri, M.; Todoerti, K.; Agnelli, L.; Marzorati, S.; Fabris, S.; Ciceri, G.; Galletti, S.; Milesi, G.; Manzoni, M.; et al. Molecular Spectrum of BRAF, NRAS and KRAS Gene Mutations in Plasma Cell Dyscrasias: Implication for MEK-ERK Pathway Activation. Oncotarget 2015, 6, 24205–24217. [Google Scholar] [CrossRef]
  63. Xu, J.; Pfarr, N.; Endris, V.; Mai, E.K.; Md Hanafiah, N.H.; Lehners, N.; Penzel, R.; Weichert, W.; Ho, A.D.; Schirmacher, P.; et al. Molecular Signaling in Multiple Myeloma: Association of RAS/RAF Mutations and MEK/ERK Pathway Activation. Oncogenesis 2017, 6, e337. [Google Scholar] [CrossRef] [PubMed]
  64. Mey, U.J.M.; Renner, C.; von Moos, R. Vemurafenib in Combination with Cobimetinib in Relapsed and Refractory Extramedullary Multiple Myeloma Harboring the BRAF V600E Mutation. Hematol. Oncol. 2017, 35, 890–893. [Google Scholar] [CrossRef] [PubMed]
  65. Hofste Op Bruinink, D.; Hoogenboezem, R.; Bindels, E.; Sanders, M.; Erpelinck—Verschueren, C.; Van Strien, P.; Koenders, J.; Misund, K.; Beverloo, B.; Van Der Holt, B.; et al. Abstract 5023: Multiple Myeloma with a Clonal Del17p Aberration Is Characterized by Somatic TP53 Mutations, Which Negatively Affect Prognosis in This Cytogenetic Subgroup. Cancer Res. 2016, 76, 5023. [Google Scholar] [CrossRef]
  66. Xiong, W.; Wu, X.; Starnes, S.; Johnson, S.K.; Haessler, J.; Wang, S.; Chen, L.; Barlogie, B.; Shaughnessy, J.D.; Zhan, F. An Analysis of the Clinical and Biologic Significance of TP53 Loss and the Identification of Potential Novel Transcriptional Targets of TP53 in Multiple Myeloma. Blood 2008, 112, 4235–4246. [Google Scholar] [CrossRef]
  67. Petrilla, C.; Galloway, J.; Kudalkar, R.; Ismael, A.; Cottini, F. Understanding DNA Damage Response and DNA Repair in Multiple Myeloma. Cancers 2023, 15, 4155. [Google Scholar] [CrossRef]
  68. 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]
  69. Shyamsunder, P.; Sridharan, S.P.; Madan, V.; Dakle, P.; Zeya, C.; Kanojia, D.; Chng, W.-J.; Ong, S.T.; Koeffler, H.P. THZ531 Induces a State of BRCAness in Multiple Myeloma Cells: Synthetic Lethality with Combination Treatment of THZ 531 with DNA Repair Inhibitors. IJMS 2022, 23, 1207. [Google Scholar] [CrossRef]
  70. Rahmat, M.; Clement, K.; Sklavenitis-Pistofidis, R.; Kodgule, R.; Fulco, C.; Alberge, J.-B.; Boehner, C.J.; Agius, M.P.; Kitzenberg, E.M.; Dorfman, D.; et al. Identification of a Novel Epigenetic Mechanism of MYC Deregulation in Smoldering and Newly Diagnosed Multiple Myeloma Patients. Blood 2021, 138, 504. [Google Scholar] [CrossRef]
  71. 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, 1055–1066. [Google Scholar] [CrossRef]
  72. Walker, B.A.; Wardell, C.P.; Begum, D.; Dahir, N.; Johnson, D.C.; Davies, F.E.; Morgan, G.J. MYC Translocations In Multiple Myeloma Involve Recruitment Of Enhancer Elements Resulting In Over-Expression and Decreased Overall Survival. Blood 2013, 122, 274. [Google Scholar] [CrossRef]
  73. Rahmat, M.; Clement, K.; Alberge, J.-B.; Sklavenitis-Pistofidis, R.; Kodgule, R.; Fulco, C.P.; Heilpern-Mallory, D.; Nilsson, K.; Dorfman, D.; Engreitz, J.M.; et al. Novel Mechanism of MYCD eregulation in Multiple Myeloma. bioRxiv 2023. [Google Scholar] [CrossRef]
  74. Handa, H.; Honma, K.; Oda, T.; Kobayashi, N.; Kuroda, Y.; Kimura-Masuda, K.; Watanabe, S.; Ishihara, R.; Murakami, Y.; Masuda, Y.; et al. Long Noncoding RNA PVT1 Is Regulated by Bromodomain Protein BRD4 in Multiple Myeloma and Is Associated with Disease Progression. IJMS 2020, 21, 7121. [Google Scholar] [CrossRef]
  75. Anguiano, A.; Acharya, C.; Salter, K.; McCluskey, D.; Gasperetto, C.; Zhan, F.; Dhodapkar, M.; Nevins, J.; Barlogie, B.; Shaughnessy, J.D.; et al. Gene Expression Profiles for Prognosis in MGUS, Coupled with Signatures of Oncogenic Pathway Deregulation Provide a Novel Approach for Selection of Molecular Targets in Multiple Myeloma. Blood 2007, 110, 655. [Google Scholar] [CrossRef]
  76. Alderton, G.K. Targeting MYC? You BET. Nat. Rev. Drug Discov. 2011, 10, 732–733. [Google Scholar] [CrossRef] [PubMed]
  77. 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]
  78. Maura, F.; Bolli, N.; Minvielle, S.; Gloznik, D.; Szalat, R.; Fullam, A.; Martincorena, I.; Samur, M.K.; Tarpey, P.; Davies, H.; et al. Analysis of Mutational Signatures Suggest That Aid Has an Early and Driver Role in Multiple Myeloma. Blood 2016, 128, 116. [Google Scholar] [CrossRef]
  79. Farswan, A.; Gupta, A.; Jena, L.; Ruhela, V.; Kaur, G.; Gupta, R. Characterizing the Mutational Landscape of MM and Its Precursor MGUS. Am. J. Cancer Res. 2022, 12, 1919–1933. [Google Scholar]
  80. Aktas Samur, A.; Fulciniti, M.; Szalat, R.; Corre, J.; Bolli, N.; Moreau, P.; Attal, M.; Anderson, K.C.; Parmigiani, G.; Avet-Loiseau, H.; et al. Lack of Significant Differences in Somatic Alterations between MGUS, SMM and Symptomatic Multiple Myeloma: A Result from Comprehensive Genomic Profiling Study. Blood 2019, 134, 3089. [Google Scholar] [CrossRef]
  81. Ohguchi, H.; Hideshima, T.; Anderson, K.C. The Biological Significance of Histone Modifiers in Multiple Myeloma: Clinical Applications. Blood Cancer J. 2018, 8, 83. [Google Scholar] [CrossRef]
  82. Walker, B.A.; Leone, P.E.; Dickens, N.J.; Boyd, K.D.; Gonzalez, D.; Davies, F.E.; Morgan, G.J. UTX, a Histone Demethylase, Is Inactivated through Homozygous Deletion, Mutation, and DNA Methylation in Multiple Myeloma. Blood 2009, 114, 1798. [Google Scholar] [CrossRef]
  83. Deligezer, U.; Akisik, E.E.; Erten, N.; Dalay, N. Sequence-Specific Histone Methylation Is Detectable on Circulating Nucleosomes in Plasma. Clin. Chem. 2008, 54, 1125–1131. [Google Scholar] [CrossRef]
  84. Muers, M. Mapping Histone Modifications and DNA Methylation Together. Nat. Rev. Genet. 2012, 13, 299. [Google Scholar] [CrossRef]
  85. Fuks, F. DNA Methylation and Histone Modifications: Teaming up to Silence Genes. Curr. Opin. Genet. Dev. 2005, 15, 490–495. [Google Scholar] [CrossRef]
  86. Pawlyn, C.; Bright, M.D.; Buros, A.F.; Stein, C.K.; Walters, Z.; Aronson, L.I.; Mirabella, F.; Jones, J.R.; Kaiser, M.F.; Walker, B.A.; et al. Overexpression of EZH2 in Multiple Myeloma Is Associated with Poor Prognosis and Dysregulation of Cell Cycle Control. Blood Cancer J. 2017, 7, e549. [Google Scholar] [CrossRef] [PubMed]
  87. Harding, T.; Swanson, J.; Van Ness, B. EZH2 Inhibitors Sensitize Myeloma Cell Lines to Panobinostat Resulting in Unique Combinatorial Transcriptomic Changes. Oncotarget 2018, 9, 21930–21942. [Google Scholar] [CrossRef]
  88. Adamik, J.; Jin, S.; Sun, Q.; Zhang, P.; Weiss, K.R.; Anderson, J.L.; Silbermann, R.; Roodman, G.D.; Galson, D.L. EZH2 or HDAC1 Inhibition Reverses Multiple Myeloma–Induced Epigenetic Suppression of Osteoblast Differentiation. Mol. Cancer Res. 2017, 15, 405–417. [Google Scholar] [CrossRef]
  89. Carew, J.S.; Espitia, C.M.; Zhao, W.; Visconte, V.; Anwer, F.; Kelly, K.R.; Nawrocki, S.T. Rational Cotargeting of HDAC6 and BET Proteins Yields Synergistic Antimyeloma Activity. Blood Adv. 2019, 3, 1318–1329. [Google Scholar] [CrossRef]
  90. Alzrigat, M.; Párraga, A.A.; Majumder, M.M.; Ma, A.; Jin, J.; Österborg, A.; Nahi, H.; Nilsson, K.; Heckman, C.A.; Öberg, F.; et al. The Polycomb Group Protein BMI-1 Inhibitor PTC-209 Is a Potent Anti-Myeloma Agent Alone or in Combination with Epigenetic Inhibitors Targeting EZH2 and the BET Bromodomains. Oncotarget 2017, 8, 103731–103743. [Google Scholar] [CrossRef]
  91. Puła, A.; Robak, P.; Robak, T. MicroRNA in Multiple Myeloma—A Role in Pathogenesis and Prognostic Significance. CMC 2021, 28, 6753–6772. [Google Scholar] [CrossRef]
  92. Zhang, R.; Zhang, D.; Luo, Y.; Sun, Y.; Duan, C.; Yang, J.; Wei, J.; Li, X.; Lu, Y.; Lai, X. miR-34a Promotes the Immunosuppressive Function of Multiple Myeloma-Associated Macrophages by Dampening the TLR-9 Signaling. Cancer Med. 2024, 13, e7387. [Google Scholar] [CrossRef]
  93. Szudy-Szczyrek, A.; Ahern, S.; Krawczyk, J.; Szczyrek, M.; Hus, M. MiRNA as a Potential Target for Multiple Myeloma Therapy–Current Knowledge and Perspectives. J. Pers. Med. 2022, 12, 1428. [Google Scholar] [CrossRef] [PubMed]
  94. Zhang, Y.-K.; Wang, H.; Leng, Y.; Li, Z.-L.; Yang, Y.-F.; Xiao, F.-J.; Li, Q.-F.; Chen, X.-Q.; Wang, L.-S. Overexpression of microRNA-29b Induces Apoptosis of Multiple Myeloma Cells through down Regulating Mcl-1. Biochem. Biophys. Res. Commun. 2011, 414, 233–239. [Google Scholar] [CrossRef] [PubMed]
  95. Long, S.; Long, S.; He, H.; Chen, G. microRNA-765 Is Pregulated in Multiple Myeloma and Serves an Oncogenic Role by Directly Targeting SOX6. Exp. Ther. Med. 2019, 17, 4741–4747. [Google Scholar] [CrossRef] [PubMed]
  96. Kimura, K.; Kuroda, Y.; Masuda, Y.; Yamane, A.; Hattori, H.; Tahara, K.; Kaneko, A.; Suda, I.; Takahashi, N.; Gotoh, N.; et al. Loop Regulation Between microRNAs and Epigenetics Underlie microRNA Dysregulation in Multiple Myeloma and Is Associated with the Disease Progression. Blood 2015, 126, 3013. [Google Scholar] [CrossRef]
  97. Gregorova, J.; Vlachova, M.; Vychytilova-Faltejskova, P.; Dostalova, A.; Ruzickova, T.; Vecera, M.; Radova, L.; Pospichalova, V.; Sladecek, S.; Hyzdalova, M.; et al. MicroRNA Profiling of Bone Marrow Plasma Extracellular Vesicles in Multiple Myeloma, Extramedullary Disease, and Plasma Cell Leukemia. Hematol. Oncol. 2025, 43, e70036. [Google Scholar] [CrossRef]
  98. Rossi, M.; Tagliaferri, P.; Tassone, P. MicroRNAs in Multiple Myeloma and Related Bone Disease. Ann. Transl. Med. 2015, 3, 334. [Google Scholar] [CrossRef]
  99. Di Martino, M.T.; Leone, E.; Amodio, N.; Foresta, U.; Lionetti, M.; Pitari, M.R.; Gallo Cantafio, M.E.; Gullà, A.; Conforti, F.; Morelli, E.; et al. Synthetic miR-34a Mimics as a Novel Therapeutic Agent for Multiple Myeloma: In Vitro and in Vivo Evidence. Clin. Cancer Res. 2012, 18, 6260–6270. [Google Scholar] [CrossRef]
  100. Zarone, M.R.; Misso, G.; Grimaldi, A.; Zappavigna, S.; Russo, M.; Amler, E.; Di Martino, M.T.; Amodio, N.; Tagliaferri, P.; Tassone, P.; et al. Evidence of Novel miR-34a-Based Therapeutic Approaches for Multiple Myeloma Treatment. Sci. Rep. 2017, 7, 17949. [Google Scholar] [CrossRef]
  101. Kastritis, E.; Malandrakis, P.; Solia, I.; Ntanasis-Stathopoulos, I.; Theodorakakou, F.; Spiliopoulou, V.; Fotiou, D.; Migkou, M.; Kanellias, N.; Roussou, M.; et al. Risk and Patterns of Progression Among Smoldering Myeloma Patients after the Implementation of New Criteria and Modern Imaging. Blood 2024, 144, 4675. [Google Scholar] [CrossRef]
  102. Rajkumar, S.V. Evolving Diagnostic Criteria for Multiple Myeloma. Hematology 2015, 2015, 272–278. [Google Scholar] [CrossRef] [PubMed]
  103. Ji, M.; Wang, M.; Shih, Y.-H.; Huber, J.; Fiala, M.A.; Wang, R.; Sanfilippo, K.M.; Thomas, T.S.; Wang, S.-Y.; Schoen, M.W.; et al. From Criteria to Clinic: How Updated SLiM CRAB Criteria Influence Multiple Myeloma Diagnostic Activity. JCO 2024, 42, 7556. [Google Scholar] [CrossRef]
  104. Ludwig, H.; Kainz, S.; Schreder, M.; Zojer, N.; Hinke, A. SLiM CRAB Criteria Revisited: Temporal Trends in Prognosis of Patients with Smoldering Multiple Myeloma Who Meet the Definition of ‘Biomarker-Defined Early Multiple Myeloma’—A Systematic Review with Meta-Analysis. EClinicalMedicine 2023, 58, 101910. [Google Scholar] [CrossRef] [PubMed]
  105. Kumar, S.K.; Uno, H.; Jacobus, S.J.; Wier, S.A.V.; Ahmann, G.J.; Henderson, K.J.; Callander, N.S.; Haug, J.L.; Siegel, D.S.; Greipp, P.R.; et al. Impact of Gene Expression Profiling-Based Risk Stratification in Patients with Myeloma Receiving Initial Therapy with Lenalidomide and Dexamethasone. Blood 2011, 118, 4359–4362. [Google Scholar] [CrossRef]
  106. Fonseca, R.; Braggio, E.; Keats, J.J.; Ahmann, G.J.; Mantei, J.; Stewart, A.K.; Bergsagel, P.L. Generation of an Automated Tool for Querying Myeloma Transcriptomics for Multiple Gene Expression Signatures Used in Risk-Stratification. JCO 2011, 29, 8024. [Google Scholar] [CrossRef]
  107. Meißner, T.; Seckinger, A.; Rème, T.; Hielscher, T.; Möhler, T.; Neben, K.; Goldschmidt, H.; Klein, B.; Hose, D. Metascoring and Gene Expression Profiling in Clinical Routine in Multiple Myeloma. Blood 2011, 118, 3940. [Google Scholar] [CrossRef]
  108. Heuck, C.J.; Johnson, S.K.; Zhang, Q.; Shaughnessy, J.D. Gene Expression Signature in MGUS and Multiple Myeloma. In Genetic and Molecular Epidemiology of Multiple Myeloma; Lentzsch, S., Ed.; Springer: New York, NY, USA, 2013; pp. 17–41. ISBN 978-1-4614-4660-6. [Google Scholar]
  109. Cerchione, C.; Usmani, S.Z.; Stewart, A.K.; Kaiser, M.; Rasche, L.; Kortüm, M.; Mateos, M.-V.; Spencer, A.; Sonneveld, P.; Anderson, K.C. Gene Expression Profiling in Multiple Myeloma: Redefining the Paradigm of Risk-Adapted Treatment. Front. Oncol. 2022, 12, 820768. [Google Scholar] [CrossRef]
  110. D’Agostino, M.; Cairns, D.A.; Lahuerta, J.J.; Wester, R.; Bertsch, U.; Waage, A.; Zamagni, E.; Mateos, M.-V.; Dall’Olio, D.; van de Donk, N.W.C.J.; et al. Second Revision of the International Staging System (R2-ISS) for Overall Survival in Multiple Myeloma: A European Myeloma Network (EMN) Report Within the HARMONY Project. JCO 2022, 40, 3406–3418. [Google Scholar] [CrossRef]
  111. D’Agostino, M.; Zaccaria, G.M.; Ziccheddu, B.; Rustad, E.H.; Genuardi, E.; Capra, A.; Oliva, S.; Auclair, D.; Yesil, J.; Colucci, P.; et al. Early Relapse Risk in Patients with Newly Diagnosed Multiple Myeloma Characterized by Next-Generation Sequencing. Clin. Cancer Res. 2020, 26, 4832–4841. [Google Scholar] [CrossRef]
  112. Fiala, M.A.; Dukeman, J.; Stockerl-Goldstein, K.; Tomasson, M.H.; Wildes, T.M.; Auclair, D.; Vij, R. Next Generation Sequencing Based Revised International Staging System (R-ISS) for Multiple Myeloma. Blood 2016, 128, 2349. [Google Scholar] [CrossRef]
  113. Szalat, R.; Munshi, N.C. Next-Generation Sequencing Informing Therapeutic Decisions and Personalized Approaches. Am. Soc. Clin. Oncol. Educ. Book 2016, 36, e442–e448. [Google Scholar] [CrossRef] [PubMed]
  114. Pratiwi, L.; Mashudi, F.H.; Ningtyas, M.C.; Sutanto, H.; Romadhon, P.Z. Genetic Profiling of Acute and Chronic Leukemia via Next-Generation Sequencing: Current Insights and Future Perspectives. Hematol. Rep. 2025, 17, 18. [Google Scholar] [CrossRef] [PubMed]
  115. Sutanto, H.; Pratiwi, L.; Romadhon, P.Z.; Bintoro, S.U.Y. Advancing Chronic Myeloid Leukemia Research with Next-Generation Sequencing: Potential Benefits, Limitations, and Future Clinical Integration. Hum. Genet. 2025, 144, 481–503. [Google Scholar] [CrossRef] [PubMed]
  116. Lagana, A.; Perumal, D.; Melnekoff, D.; Readhead, B.; Kidd, B.; Leshchenko, V.V.; Kuo, P.-Y.; Keats, J.J.; Derome, M.; Yesil, J.; et al. Integrative Network Analysis of Newly Diagnosed Multiple Myeloma Identifies a Novel RNA-Seq Based High Riskgene Signature. Blood 2016, 128, 3285. [Google Scholar] [CrossRef]
  117. 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]
  118. Gozzetti, A.; Bocchia, M. Minimal Residual Disease in Multiple Myeloma: An Important Tool inClinical Trials. RRCT 2022, 17, 9–10. [Google Scholar] [CrossRef]
  119. Ferla, V.; Antonini, E.; Perini, T.; Farina, F.; Masottini, S.; Malato, S.; Marktel, S.; Lupo Stanghellini, M.T.; Tresoldi, C.; Ciceri, F.; et al. Minimal Residual Disease Detection by Next-Generation Sequencing in Multiple Myeloma: Promise and Challenges for Response-Adapted Therapy. Front. Oncol. 2022, 12, 932852. [Google Scholar] [CrossRef]
  120. Mina, R.; Bonello, F.; Oliva, S. Minimal Residual Disease in Multiple Myeloma: Ready for Prime Time? Cancer J. 2021, 27, 247–255. [Google Scholar] [CrossRef]
  121. Aljama, M.A.; Sidiqi, H.M.; Gertz, M.A. Are We Maintaining Minimal Residual Disease in Myeloma? Leuk. Lymphoma 2025, 66, 1001–1009. [Google Scholar] [CrossRef]
  122. Yao, Q.; Bai, Y.; Orfao, A.; Kumar, S.; Chim, C.S. Upgraded Standardized Minimal Residual Disease Detection by Next-Generation Sequencing in Multiple Myeloma. J. Mol. Diagn. 2020, 22, 679–684. [Google Scholar] [CrossRef]
  123. Mina, R.; Oliva, S.; Boccadoro, M. Minimal Residual Disease in Multiple Myeloma: State of the Art and Future Perspectives. JCM 2020, 9, 2142. [Google Scholar] [CrossRef]
  124. Genovesi, L.; Schirm, M.; Dupuis, N.; Pottiez, G. Measuring Minimal Residual Disease in Plasma of Multiple Myeloma Patients By Intact Mass Spectrometry. Blood 2023, 142, 6652. [Google Scholar] [CrossRef]
  125. Medina-Herrera, A.; Sarasquete, M.E.; Jiménez, C.; Puig, N.; García-Sanz, R. Minimal Residual Disease in Multiple Myeloma: Past, Present, and Future. Cancers 2023, 15, 3687. [Google Scholar] [CrossRef] [PubMed]
  126. Terpos, E.; Dimopoulos, M.A.; Moulopoulos, L.A. The Role of Imaging in the Treatment of Patients with Multiple Myeloma in 2016. Am. Soc. Clin. Oncol. Educ. Book 2016, 36, e407–e417. [Google Scholar] [CrossRef] [PubMed]
  127. Withofs, N.; Nanni, C.; Simoni, P.; Fanti, S.; Beguin, Y.; Caers, J. Imaging Myeloma and Related Monoclonal Plasma Cell Disorders Using MRI, Low-Dose Whole-Body CT and FDG PET/CT. Clin. Transl. Imaging 2015, 3, 95–109. [Google Scholar] [CrossRef]
  128. Rama, S.; Suh, C.H.; Kim, K.W.; Durieux, J.C.; Ramaiya, N.H.; Tirumani, S.H. Comparative Performance of Whole-Body MRI and FDG PET/CT in Evaluation of Multiple Myeloma Treatment Response: Systematic Review and Meta-Analysis. Am. J. Roentgenol. 2022, 218, 602–613. [Google Scholar] [CrossRef]
  129. Hillengass, J.; Moulopoulos, L.A.; Delorme, S.; Koutoulidis, V.; Mosebach, J.; Hielscher, T.; Drake, M.; Rajkumar, S.V.; Oestergaard, B.; Abildgaard, N.; et al. Whole-Body Computed Tomography versus Conventional Skeletal Survey in Patients with Multiple Myeloma: A Study of the International Myeloma Working Group. Blood Cancer J. 2017, 7, e599. [Google Scholar] [CrossRef]
  130. Pianko, M.J.; Terpos, E.; Roodman, G.D.; Divgi, C.R.; Zweegman, S.; Hillengass, J.; Lentzsch, S. Whole-Body Low-Dose Computed Tomography and Advanced Imaging Techniques for Multiple Myeloma Bone Disease. Clin. Cancer Res. 2014, 20, 5888–5897. [Google Scholar] [CrossRef]
  131. Prieto, E.; García-Velloso, M.J.; Aquerreta, J.D.; Rosales, J.J.; Bastidas, J.F.; Soriano, I.; Irazola, L.; Rodríguez-Otero, P.; Quincoces, G.; Martí-Climent, J.M. Ultra-Low Dose Whole-Body CT for Attenuation Correction in a Dual Tracer PET/CT Protocol for Multiple Myeloma. Phys. Medica 2021, 84, 1–9. [Google Scholar] [CrossRef]
  132. Wu, F.; Bernard, S.; Fayad, L.M.; Ilaslan, H.; Messiou, C.; Moulopoulos, L.A.; Mulligan, M.E. Updates and Ongoing Challenges in Imaging of Multiple Myeloma: AJR Expert Panel Narrative Review. Am. J. Roentgenol. 2021, 217, 775–785. [Google Scholar] [CrossRef]
  133. Mosebach, J.; Thierjung, H.; Schlemmer, H.-P.; Delorme, S. Multiple Myeloma Guidelines and Their Recent Updates: Implications for Imaging. Fortschr. Röntgenstr. 2019, 191, 998–1009. [Google Scholar] [CrossRef]
  134. 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] [PubMed]
  135. Sonneveld, P.; Dimopoulos, M.A.; Boccadoro, M.; Quach, H.; Ho, P.J.; Beksac, M.; Hulin, C.; Antonioli, E.; Leleu, X.; Mangiacavalli, S.; et al. Daratumumab, Bortezomib, Lenalidomide, and Dexamethasone for Multiple Myeloma. N. Engl. J. Med. 2024, 390, 301–313. [Google Scholar] [CrossRef]
  136. Usmani, S.Z.; Facon, T.; Hungria, V.; Bahlis, N.J.; Venner, C.P.; Braunstein, M.; Pour, L.; Martí, J.M.; Basu, S.; Cohen, Y.C.; et al. Daratumumab plus Bortezomib, Lenalidomide and Dexamethasone for Transplant-Ineligible or Transplant-Deferred Newly Diagnosed Multiple Myeloma: The Randomized Phase 3 CEPHEUS Trial. Nat. Med. 2025, 31, 1195–1202. [Google Scholar] [CrossRef] [PubMed]
  137. Hansen, D.K.; Gautam, S.; Lafeuille, M.; Rossi, C.; Moore, B.; Tardif-Samson, A.; Thompson-Leduc, P.; Fu, A.Z.; Cortoos, A.; Kaila, S.; et al. Comparison of Time to Next Treatment or Death Between Front-Line Daratumumab, Lenalidomide, and Dexamethasone (DRd) Versus Bortezomib, Lenalidomide, and Dexamethasone (VRd) Among Transplant-Ineligible Patients with Multiple Myeloma. Cancer Med. 2024, 13, e70308. [Google Scholar] [CrossRef] [PubMed]
  138. Byun, J.M.; Park, S.-S.; Yoon, S.-S.; Ahn, A.; Kim, M.; Lee, J.Y.; Jeon, Y.-W.; Shin, S.-H.; Yahng, S.-A.; Koh, Y.; et al. Advantage of Achieving Deep Response Following Frontline Daratumumab-VTd Compared to VRd in Transplant-Eligible Multiple Myeloma: Multicenter Study. Blood Res. 2023, 58, 83–90. [Google Scholar] [CrossRef] [PubMed]
  139. Garrido, D.; Bove, V.; Villano, F.; Riva, E. Survival Analysis of Newly Diagnosed Multiple Myeloma Patients after Frontline Autologous Stem Cell Transplantation in a Real-Life Setting. Acta Med. 2023, 66, 117–121. [Google Scholar] [CrossRef]
  140. Pasvolsky, O.; Wang, Z.; Milton, D.R.; Tanner, M.R.; Bashir, Q.; Srour, S.; Saini, N.; Lin, P.; Ramdial, J.; Nieto, Y.; et al. Multiple Myeloma Patients with a Long Remission after Autologous Hematopoietic Stem Cell Transplantation. Blood Cancer J. 2024, 14, 82. [Google Scholar] [CrossRef]
  141. Pawlyn, C.; Cairns, D.; Menzies, T.; Jones, J.; Jenner, M.; Cook, G.; Boyd, K.; Drayson, M.; Kaiser, M.; Owen, R.; et al. Autologous Stem Cell Transplantation Is Safe and Effective for Fit, Older Myeloma Patients: Exploratory Results from the Myeloma XI Trial. Haematologica 2020, 107, 231–242. [Google Scholar] [CrossRef]
  142. Rosenberg, A.S.; Brunson, A.; Jonas, B.A.; Keegan, T.H.M.; Wun, T. Association Between Autologous Stem Cell Transplant and Survival Among Californians with Multiple Myeloma. JNCI J. Natl. Cancer Inst. 2019, 111, 78–85. [Google Scholar] [CrossRef]
  143. Ricciuti, G.; Falcone, A.; Cascavilla, N.; Martinelli, G.; Cerchione, C. Autologous Stem Cell Transplantation in Multiple Myeloma. Panminerva Med. 2021, 62, 220–224. [Google Scholar] [CrossRef]
  144. Attal, M.; cances Lauwers, V.; Marit, G.; Caillot, D.; Facon, T.; Hulin, C.; Moreau, P.; Mathiot, C.; Roussel, M.; Payen, C.; et al. Maintenance Treatment with Lenalidomide After Transplantation for MYELOMA: Final Analysis of the IFM 2005-02. Blood 2010, 116, 310. [Google Scholar] [CrossRef]
  145. 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] [PubMed]
  146. Holstein, S.A.; Jung, S.-H.; Richardson, P.G.; Hofmeister, C.C.; Hurd, D.D.; Hassoun, H.; Giralt, S.; Stadtmauer, E.A.; Weisdorf, D.J.; Vij, R.; et al. Updated Analysis of CALGB 100104 (Alliance): A Randomised Phase III Study Evaluating Lenalidomide vs. Placebo Maintenance after Single Autologous Stem Cell Transplant for Multiple Myeloma. Lancet Haematol. 2017, 4, e431–e442. [Google Scholar] [CrossRef]
  147. Syed, Y.Y. Lenalidomide: A Review in Newly Diagnosed Multiple Myeloma as Maintenance Therapy After ASCT. Drugs 2017, 77, 1473–1480. [Google Scholar] [CrossRef]
  148. Rifkin, R.M.; Jagannath, S.; Durie, B.G.M.; Narang, M.; Terebelo, H.R.; Gasparetto, C.J.; Toomey, K.; Hardin, J.W.; Wagner, L.; Parikh, K.; et al. Treatment Outcomes and Health Care Resource Utilization in Patients with Newly Diagnosed Multiple Myeloma Receiving Lenalidomide-Only Maintenance, Any Maintenance, or No Maintenance: Results from the Connect MM Registry. Clin. Ther. 2018, 40, 1193–1202.e1. [Google Scholar] [CrossRef]
  149. Alonso, R.; Cedena, M.-T.; Wong, S.; Shah, N.; Ríos-Tamayo, R.; Moraleda, J.M.; López-Jiménez, J.; García, C.; Bahri, N.; Valeri, A.; et al. Prolonged Lenalidomide Maintenance Therapy Improves the Depth of Response in Multiple Myeloma. Blood Adv. 2020, 4, 2163–2171. [Google Scholar] [CrossRef]
  150. 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]
  151. Yamamoto, M.; Ohashi, K.; Kakihana, K.; Nakamura, Y.; Komeno, T.; Kojima, H.; Morita, S.; Sakamaki, H. A Multicenter Clinical Study to Determine the Feasible Initial Dose of Lenalidomide for Maintenance Therapy in Patients with Multiple Myeloma Following Autologous Peripheral Blood Stem-Cell Transplantation. Mol. Clin. Oncol. 2016, 4, 965–970. [Google Scholar] [CrossRef]
  152. Parrondo, R.D.; Reljic, T.; Iqbal, M.; Ayala, E.; Kharfan-Dabaja, M.A.; Kumar, A.; Murthy, H. Efficacy of Proteasome Inhibitor-Based Maintenance Post Autologous Hematopoietic Cell Transplant in Multiple Myeloma: A Systematic Review and Meta-Analysis. Biol. Blood Marrow Transplant. 2020, 26, S244–S245. [Google Scholar] [CrossRef]
  153. Karam, D.; Gertz, M.A.; Lacy, M.Q.; Dispenzieri, A.; Hayman, S.R.; Dingli, D.; Kapoor, P.; Kourelis, T.; Warsame, R.M.; Hogan, W.J.; et al. Use of Maintenance Therapy Post Autologous Stem Cell Transplantation Outside of Clinical Trial Setting for Multiple Myeloma: Single Institution Experience. Blood 2019, 134, 2013. [Google Scholar] [CrossRef]
  154. McCarthy, P.L.; Owzar, K.; Hahn, T. Autologous Hematopoietic Stem Cell Transplantation and Maintenance Therapy for Multiple Myeloma. Int. J. Hematol. Oncol. 2013, 2, 71–83. [Google Scholar] [CrossRef]
  155. Sonneveld, P.; Schmidt-Wolf, I.; van der Holt, B.; el Jarari, L.; Bertsch, U.; Salwender, H.; Zweegman, S.; Vellenga, E.; Schubert, J.; Blau, I.W.; et al. HOVON-65/GMMG-HD4 Randomized Phase III Trial Comparing Bortezomib, Doxorubicin, Dexamethasone (PAD) Vs. VAD Followed by High-Dose Melphalan (HDM) and Maintenance with Bortezomib or Thalidomide In Patients with Newly Diagnosed Multiple Myeloma (MM). Blood 2010, 116, 40. [Google Scholar] [CrossRef]
  156. 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. JCO 2012, 30, 2946–2955. [Google Scholar] [CrossRef]
  157. 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]
  158. Dimopoulos, M.A.; Špička, I.; Quach, H.; Oriol, A.; Hájek, R.; Garg, M.; Beksac, M.; Bringhen, S.; Katodritou, E.; Chng, W.-J.; et al. Ixazomib as Postinduction Maintenance for Patients with Newly Diagnosed Multiple Myeloma Not Undergoing Autologous Stem Cell Transplantation: The Phase III TOURMALINE-MM4 Trial. JCO 2020, 38, 4030–4041. [Google Scholar] [CrossRef]
  159. Martello, M.; Poletti, A.; Borsi, E.; Solli, V.; Dozza, L.; Barbato, S.; Zamagni, E.; Tacchetti, P.; Pantani, L.; Mancuso, K.; et al. Clonal and Subclonal TP53 Molecular Impairment Is Associated with Prognosis and Progression in Multiple Myeloma. Blood Cancer J. 2022, 12, 15. [Google Scholar] [CrossRef] [PubMed]
  160. 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]
  161. Novoa Jáuregui, S.; Gabarros Subira, M.; Garrido, A.M.; Palomo, L.; Medina, I.; Beas, F.; Bosch, M.; Muzio, S.; Camarillas, S.; Hidalgo-Gomez, G.; et al. Ultra-Deep Sequencing of TP53 in Multiple Myeloma: Double-Hit Prevalence and Clinical Phenotype. Blood 2024, 144, 1935. [Google Scholar] [CrossRef]
  162. Munawar, U.; Barrio, S.; Roth, M.; Einsele, H.; Bargou, R.C.; Stuehmer, T.; Kortum, M. Implications of TP53 Alterations for Therapy Response in Multiple Myeloma. Blood 2018, 132, 3189. [Google Scholar] [CrossRef]
  163. Munawar, U.; Roth, M.; Barrio, S.; Wajant, H.; Siegmund, D.; Bargou, R.C.; Kortüm, K.M.; Stühmer, T. Assessment of TP53 Lesions for P53 System Functionality and Drug Resistance in Multiple Myeloma Using an Isogenic Cell Line Model. Sci. Rep. 2019, 9, 18062. [Google Scholar] [CrossRef]
  164. Singh, C.; Sreedharanunni, S.; Panakkal, V.; Jandial, A.; Jain, A.; Lad, D.; Prakash, G.; Khadwal, A.; Malhotra, P. Early Mortality and Aggressive Presentation Lead to Poor Outcomes in Patients with Newly Diagnosed Double Hit and Triple Hit Multiple Myeloma. Blood 2021, 138, 3762. [Google Scholar] [CrossRef]
  165. D’Agostino, M.; De Paoli, L.; Conticello, C.; Offidani, M.; Ria, R.; Petrucci, M.T.; Spada, S.; Marcatti, M.; Catalano, L.; Gilestro, M.; et al. Continuous Therapy in Standard- and High-Risk Newly-Diagnosed Multiple Myeloma: A Pooled Analysis of 2 Phase III Trials. Crit. Rev. Oncol. Hematol. 2018, 132, 9–16. [Google Scholar] [CrossRef] [PubMed]
  166. Marneni, N.; Chakraborty, R. Current Approach to Managing Patients with Newly Diagnosed High-Risk Multiple Myeloma. Curr. Hematol. Malig. Rep. 2021, 16, 148–161. [Google Scholar] [CrossRef]
  167. Cazaubiel, T.; Mulas, O.; Montes, L.; Schavgoulidze, A.; Avet-Loiseau, H.; Corre, J.; Perrot, A. Risk and Response-Adapted Treatment in Multiple Myeloma. Cancers 2020, 12, 3497. [Google Scholar] [CrossRef]
  168. Goel, U.; Usmani, S.; Kumar, S. Current Approaches to Management of Newly Diagnosed Multiple Myeloma. Am. J. Hematol. 2022, 97, S3–S25. [Google Scholar] [CrossRef]
  169. Chatziravdeli, V.; Katsaras, G.N.; Katsaras, D.; Doxani, C.; Stefanidis, I.; Zintzaras, E. A Systematic Review and Meta-Analysis of Interventional Studies of Bisphosphonates and Denosumab in Multiple Myeloma and Future Perspectives. J. Musculoskelet. Neuronal Interact. 2022, 22, 596–621. [Google Scholar]
  170. Adam, Z.; Straub, J.; Krejčí, M.; Pour, L.; Brančíková, D.; Ostřížková, L.; Sandecká, V.; Štork, M. Osteoprotective therapy with bisphosphonates or denosumab in patients with multiple myeloma: Benefit and risks. Vnitr. Lek. 2017, 63, 311–321. [Google Scholar] [CrossRef]
  171. Lei, M.M.; Tavares, E.; Buzgo, E.; Lou, U.; Raje, N.; Yee, A.J. Denosumab versus Intravenous Bisphosphonate Use for Hypercalcemia in Multiple Myeloma. Leuk. Lymphoma 2022, 63, 3249–3252. [Google Scholar] [CrossRef]
  172. John, P.; Khalil, A.; Gao, J.; Nishioka, J.; Kamangar, F.; Badros, A. Effects of Denosumab Therapy on Renal Function in Multiple Myeloma Patients: A Single Center Retrospective Analysis. Blood 2024, 144, 6972. [Google Scholar] [CrossRef]
  173. Leleu, X.; Martin, T.; Weisel, K.; Schjesvold, F.; Iida, S.; Malavasi, F.; Manier, S.; Min, C.-K.; Ocio, E.M.; Pawlyn, C.; et al. Anti-CD38 Antibody Therapy for Patients with Relapsed/Refractory Multiple Myeloma: Differential Mechanisms of Action and Recent Clinical Trial Outcomes. Ann. Hematol. 2022, 101, 2123–2137. [Google Scholar] [CrossRef]
  174. Moreno, L.; Perez, C.; Zabaleta, A.; Manrique, I.; Alignani, D.; Ajona, D.; Blanco, L.; Lasa, M.; Maiso, P.; Rodriguez, I.; et al. The Mechanism of Action of the Anti-CD38 Monoclonal Antibody Isatuximab in Multiple Myeloma. Clin. Cancer Res. 2019, 25, 3176–3187. [Google Scholar] [CrossRef]
  175. Yashar, D.; Regidor, B.S.S.; Jew, S.K.; Bujarski, S.; Goldwater, M.-S.; Swift, R.; Schwartz, G.E.; Eshaghian, S.; Vescio, R.; Berenson, J.R. Retrospective Analysis of Response Rates to Anti-CD38 Monoclonal Antibody Containing Regimens Among Multiple Myeloma Patients with Prior Exposure to Daratumumab or Isatuximab. Blood 2021, 138, 3785. [Google Scholar] [CrossRef]
  176. Diamond, B.; Baughn, L.; Poorebrahim, M.; Poos, A.M.; Lee, H.; Kaddoura, M.; Wiedmeier-Nutor, J.E.; Durante, M.; Otteson, G.; Jevremovic, D.; et al. CD38 Biallelic Loss Is a Recurrent Mechanism of Resistance to Anti-CD38 Antibodies in Multiple Myeloma. bioRxiv 2024. [Google Scholar] [CrossRef]
  177. Wang, H.; Koob, T.; Fromm, J.R.; Gopal, A.; Carter, D.; Lieber, A. CD46 and CD59 Inhibitors Enhance Complement-Dependent Cytotoxicity of Anti-CD38 Monoclonal Antibodies Daratumumab and Isatuximab in Multiple Myeloma and Other B-Cell Malignancy Cells. Cancer Biol. Ther. 2024, 25, 2314322. [Google Scholar] [CrossRef] [PubMed]
  178. Einsele, H.; Schreder, M. Treatment of Multiple Myeloma with the Immunostimulatory SLAMF7 Antibody Elotuzumab. Ther. Adv. Hematol. 2016, 7, 288–301. [Google Scholar] [CrossRef] [PubMed]
  179. Pazina, T.; James, A.M.; MacFarlane, A.W.; Bezman, N.A.; Henning, K.A.; Bee, C.; Graziano, R.F.; Robbins, M.D.; Cohen, A.D.; Campbell, K.S. The Anti-SLAMF7 Antibody Elotuzumab Mediates NK Cell Activation through Both CD16-Dependent and –Independent Mechanisms. OncoImmunology 2017, 6, e1339853. [Google Scholar] [CrossRef] [PubMed]
  180. Magen, H.; Muchtar, E. Elotuzumab: The First Approved Monoclonal Antibody for Multiple Myeloma Treatment. Ther. Adv. Hematol. 2016, 7, 187–195. [Google Scholar] [CrossRef]
  181. Taniwaki, M.; Yoshida, M.; Matsumoto, Y.; Shimura, K.; Kuroda, J.; Kaneko, H. Elotuzumab for the Treatment of Relapsed or Refractory Multiple Myeloma, with Special Reference to its Modes of Action and SLAMF7 Signaling. Mediterr. J. Hematol. Infect. Dis. 2018, 10, e2018014. [Google Scholar] [CrossRef]
  182. 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]
  183. Danhof, S.; Rasche, L.; Mottok, A.; Steinmüller, T.; Zhou, X.; Schreder, M.; Kilian, T.; Strifler, S.; Rosenwald, A.; Hudecek, M.; et al. Elotuzumab for the Treatment of Extramedullary Myeloma: A Retrospective Analysis of Clinical Efficacy and SLAMF7 Expression Patterns. Ann. Hematol. 2021, 100, 1537–1546. [Google Scholar] [CrossRef]
  184. Almodovar Diaz, A.; Alouch, S.; Chawla, Y.; Gonsalves, W. The Antibody Drug Conjugate, Belantamab-Mafodotin, in the Treatment of Multiple Myeloma: A Comprehensive Review. BLCTT 2024, 14, 71–87. [Google Scholar] [CrossRef] [PubMed]
  185. Tzogani, K.; Penttilä, K.; Lähteenvuo, J.; Lapveteläinen, T.; Lopez Anglada, L.; Prieto, C.; Garcia-Ochoa, B.; Enzmann, H.; Gisselbrecht, C.; Delgado, J.; et al. EMA Review of Belantamab Mafodotin (Blenrep) for the Treatment of Adult Patients with Relapsed/Refractory Multiple Myeloma. Oncologist 2021, 26, 70–76. [Google Scholar] [CrossRef] [PubMed]
  186. 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]
  187. Nooka, A.K.; Cohen, A.D.; Lee, H.C.; Badros, A.; Suvannasankha, A.; Callander, N.; Abdallah, A.-O.; Trudel, S.; Chari, A.; Libby, E.N.; et al. Single-Agent Belantamab Mafodotin in Patients with Relapsed/Refractory Multiple Myeloma: Final Analysis of the DREAMM-2 Trial. Cancer 2023, 129, 3746–3760. [Google Scholar] [CrossRef]
  188. Abeykoon, J.P.; Vaxman, J.; Patel, S.V.; Kumar, S.; Malave, G.C.; Young, K.S.; Ailawadhi, S.; Larsen, J.T.; Dispenzieri, A.; Muchtar, E.; et al. Impact of Belantamab Mafodotin-induced Ocular Toxicity on Outcomes of Patients with Advanced Multiple Myeloma. Br. J. Haematol. 2022, 199, 95–99. [Google Scholar] [CrossRef]
  189. Terpos, E.; Gavriatopoulou, M.; Ntanasis-Stathopoulos, I.; Malandrakis, P.; Fotiou, D.; Migkou, M.; Theodorakakou, F.; Spiliopoulou, V.; Syrigou, R.; Eleutherakis-Papaiakovou, E.; et al. P884: Belantamab mafodotin plus lenalidomide and dexamethasone in transplant ineligible patients with newly diagnosed multiple myeloma: Updated results from the phase 1/2 belard study. HemaSphere 2023, 7, e828385f. [Google Scholar] [CrossRef]
  190. Lasica, M.; Spencer, A.; Campbell, P.; Wallington-Gates, C.; Wong Doo, N.; Janowski, W.; McCaughan, G.; Puliyayil, A.; Yuen, F.; Le, K.; et al. P946: A PHASE I/II SINGLE ARM STUDY OF BELANTAMAB MAFODOTIN, CARFILZOMIB AND DEXAMETHASONE IN PATIENTS WITH RELAPSED MULTIPLE MYELOMA: AMARC 19-02 BELACARD STUDY. HemaSphere 2022, 6, 836–837. [Google Scholar] [CrossRef]
  191. Eastman, S.; Shelton, C.; Gupta, I.; Krueger, J.; Blackwell, C.; Bojczuk, P.M. Synergistic Activity of Belantamab Mafodotin (Anti-BCMA Immuno-Conjugate) with PF-03084014 (Gamma-Secretase Inhibitor) in Bcma-Expressing Cancer Cell Lines. Blood 2019, 134, 4401. [Google Scholar] [CrossRef]
  192. Lonial, S.; Grosicki, S.; Hus, M.; Song, K.W.; Facon, T.; Callander, N.S.; Ribrag, V.; Uttervall, K.; Quach, H.; Vorobyev, V.I.; et al. Synergistic Effects of Low-Dose Belantamab Mafodotin in Combination with a Gamma-Secretase Inhibitor (Nirogacestat) in Patients with Relapsed/Refractory Multiple Myeloma (RRMM): DREAMM-5 Study. JCO 2022, 40, 8019. [Google Scholar] [CrossRef]
  193. Fang, J.; Zhou, F. BCMA-Targeting Chimeric Antigen Receptor T Cell Therapy for Relapsed and/or Refractory Multiple Myeloma. Ann. Hematol. 2024, 103, 1069–1083. [Google Scholar] [CrossRef]
  194. Martin, T.G.; Madduri, D.; Pacaud, L.; Usmani, S.Z. Cilta-Cel, a BCMA-Targeting CAR-T Therapy for Heavily Pretreated Patients with Relapsed/Refractory Multiple Myeloma. Future Oncol. 2023, 19, 2297–2311. [Google Scholar] [CrossRef] [PubMed]
  195. Jagannath, S.; Jackson, C.C.; Schecter, J.M.; Lendvai, N.; Sun, H.; Akram, M.; Patel, N.; Martin, T.G. Cilta-Cel, a BCMA-Targeting CAR-T Therapy for Patients with Multiple Myeloma. Expert Opin. Biol. Ther. 2024, 24, 339–350. [Google Scholar] [CrossRef]
  196. Attar, N.; Cirstea, D.; Branagan, A.R.; Yee, A.J.; Frigault, M.J.; Raje, N. Use of Cilta-Cel CAR T Cells Following Previous Use of a BCMA-Directed CAR T-Cell Product in Heavily Treated Patients with Relapsed/Refractory Multiple Myeloma: A Single Institution Case Series. Blood 2023, 142, 6924. [Google Scholar] [CrossRef]
  197. Freyer, C.W.; Porter, D.L. CAR T Toxicity Management: Cytokine Release Syndrome and Neurotoxicity. In Blood and Marrow Transplant Handbook: Comprehensive Guide for Patient Care; Maziarz, R.T., Slater, S.S., Eds.; Springer International Publishing: Cham, Switzerland, 2021; pp. 915–928. ISBN 978-3-030-53626-8. [Google Scholar]
  198. Gaffney, K.J.; Davis, J.A.; McGann, M.; Edwards, K.; Ahmed, Z.; Butcher, C.; Greenwell, B.; Hess, B.T.; Hashmi, H. Early versus Standard Management of Chimeric Antigen Receptor Therapy Toxicities and Management’s Impact on Safety and Efficacy. J. Oncol. Pharm. Pract. 2024, 30, 151–158. [Google Scholar] [CrossRef]
  199. Murthy, H.S.; Yassine, F.; Iqbal, M.; Alotaibi, S.; Moustafa, M.A.; Kharfan-Dabaja, M.A. Management of CAR T-Cell Related Toxicities: What Did the Learning Curve Teach Us so Far? Hematol. Oncol. Stem Cell Ther. 2022, 15, 100–111. [Google Scholar] [CrossRef]
  200. Walton, Z.E.; Frigault, M.J.; Maus, M.V. Current and Emerging Pharmacotherapies for Cytokine Release Syndrome, Neurotoxicity, and Hemophagocytic Lymphohistiocytosis-like Syndrome Due to CAR T Cell Therapy. Expert Opin. Pharmacother. 2024, 25, 263–279. [Google Scholar] [CrossRef]
  201. Singh, V.; Master, S. Incidence of Parkinsonism As a Complication of Anti-BCMA CAR-T Cell Therapy in Multiple Myeloma. Blood 2023, 142, 6937. [Google Scholar] [CrossRef]
  202. Graham, C.; Parker, A.; Sun, X.; Haradhvala, N.J.; Dolaher, M.; Lindell, K.; Harris, D.; Lee, W.-H.; Graham, K.J.; Nolan, H.K.; et al. Biological Correlatives of Efficacy and Toxicity of BCMA Targeting CAR T Cell Therapy in a Real-World Setting. Blood 2024, 144, 7126. [Google Scholar] [CrossRef]
  203. Gul, A.; Afaq, Y.; Khan, T.; Keen, M.A.; Ahmad, S.; Khan, M.A.; Hussain, R.; Inayat, A.; Safi, D.; Safi, S. Risk of Infection: Real-World Analysis of (Ide-Cel) Idecabtagene Vicleucel and (Cilta-Cel) Ciltacabtagene Autoleucel in Relapsed/Refractory Multiple Myeloma. Blood 2024, 144, 5160. [Google Scholar] [CrossRef]
  204. Lee, D.W.; Santomasso, B.D.; Locke, F.L.; Ghobadi, A.; Turtle, C.J.; Brudno, J.N.; Maus, M.V.; Park, J.H.; Mead, E.; Pavletic, S.; et al. ASTCT Consensus Grading for Cytokine Release Syndrome and Neurologic Toxicity Associated with Immune Effector Cells. Biol. Blood Marrow Transplant. 2019, 25, 625–638. [Google Scholar] [CrossRef]
  205. Chohan, K.L.; Siegler, E.L.; Kenderian, S.S. CAR-T Cell Therapy: The Efficacy and Toxicity Balance. Curr. Hematol. Malig. Rep. 2023, 18, 9–18. [Google Scholar] [CrossRef]
  206. Wesson, W.; Dima, D.; Suleman, N.; Saif, M.S.I.; Tabak, C.; Logan, E.; Davis, J.A.; McGann, M.; Furqan, F.; Mohan, M.; et al. Timing of Toxicities and Non-Relapse Mortality Following CAR T Therapy in Myeloma. Transplant. Cell. Ther. 2024, 30, 876–884. [Google Scholar] [CrossRef]
  207. Freyer, C.W.; Porter, D.L. Cytokine Release Syndrome and Neurotoxicity Following CAR T-Cell Therapy for Hematologic Malignancies. J. Allergy Clin. Immunol. 2020, 146, 940–948. [Google Scholar] [CrossRef] [PubMed]
  208. Rejeski, K.; Jain, M.D.; Smith, E.L. Mechanisms of Resistance and Treatment of Relapse after CAR T-Cell Therapy for Large B-Cell Lymphoma and Multiple Myeloma. Transplant. Cell. Ther. 2023, 29, 418–428. [Google Scholar] [CrossRef]
  209. Ravi, G.; Costa, L.J. Bispecific T-cell Engagers for Treatment of Multiple Myeloma. Am. J. Hematol. 2023, 98, S13–S21. [Google Scholar] [CrossRef]
  210. Verkleij, C.P.M.; Frerichs, K.A.; Broekmans, M.; Absalah, S.; Maas-Bosman, P.W.C.; Kruyswijk, S.; Nijhof, I.S.; Mutis, T.; Zweegman, S.; Van De Donk, N.W.C.J. T-Cell Redirecting Bispecific Antibodies Targeting BCMA for the Treatment of Multiple Myeloma. Oncotarget 2020, 11, 4076–4081. [Google Scholar] [CrossRef]
  211. Liu, Y.; Zhou, Y.-Q.; Nie, L.; Zhu, S.-S.; Li, N.; Wu, Z.-H.; Wang, Q.; Qi, J.; Wu, B.-Y.; Chen, S.-Q.; et al. BR109, a Novel Fully Humanized T-Cell-Engaging Bispecific Antibody with GPRC5D Binding, Has Potent Antitumor Activities in Multiple Myeloma. Cancers 2023, 15, 5774. [Google Scholar] [CrossRef]
  212. Dekhtiarenko, I.; Lelios, I.; Jacob, W.; Schneider, M.; Weisser, M.; Carlo-Stella, C.; Manier, S.; Harrison, S.J.; Popat, R.; Hasselbalch Riley, A.C.; et al. Co-Expression of GPRC5D, FcRH5 and BCMA Suggests That Targeting More Than One Cell Surface Marker May Be a Viable Strategy in Relapsed/Refractory Multiple Myeloma (RRMM): Biomarker Results from the Phase I Study of Forimtamig, a GPRC5DxCD3 Bispecific Antibody. Blood 2023, 142, 1948. [Google Scholar] [CrossRef]
  213. An, G.; Xing, L.; Chen, W.; Zhang, Y.; Gao, W.; Qiu, L.-G.; Wu, G.; Ning, J.; Wei, M.; Li, F. MBS314, a G Protein-Coupled Receptor Family C Group 5 Member D (GPRC5D) x B-Cell Maturation Antigen (BCMA) x CD3 Trispecific Antibody, in Relapsed and/or Refractory Multiple Myeloma (RRMM): Preliminary Results from a Phase I, First-in-Human, Open-Label, Dose Escalation Study. Blood 2024, 144, 3356. [Google Scholar] [CrossRef]
  214. Eckmann, J.; Fauti, T.; Biehl, M.; Zabaleta, A.; Blanco, L.; Lelios, I.; Gottwald, S.; Rae, R.; Lechner, S.; Bayer, C.; et al. Forimtamig, a Novel GPRC5D-Targeting T-Cell Bispecific Antibody with a 2+1 Format, for the Treatment of Multiple Myeloma. Blood 2025, 145, 202–219. [Google Scholar] [CrossRef]
  215. Pillarisetti, R.; Yang, D.; Yao, J.; Smith, M.; Luistro, L.; Vulfson, P.; Testa, J.; Packman, K.; Brodeur, S.; Attar, R.M.; et al. Characterization of JNJ-79635322, a Novel BCMAxGPRC5DxCD3 T-Cell Redirecting Trispecific Antibody, for the Treatment of Multiple Myeloma. Blood 2023, 142, 456. [Google Scholar] [CrossRef]
  216. Tan, Y.; Li, X.; Yu, F.; Xu, J.; Qian, Z.; Cao, Y.; Yang, X.; Du, Q.; Peng, F.; Han, S.; et al. Abstract LB128: A Novel Tri-Specific T Cell Engager Targeting BCMA and GPRC5D for Treatment of Multiple Myeloma. Cancer Res. 2024, 84, LB128. [Google Scholar] [CrossRef]
  217. Heuck, C.J.; Jethava, Y.; Khan, R.; Van Rhee, F.; Zangari, M.; Chavan, S.; Robbins, K.; Miller, S.E.; Matin, A.; Mohan, M.; et al. Inhibiting MEK in MAPK Pathway-Activated Myeloma. Leukemia 2016, 30, 976–980. [Google Scholar] [CrossRef]
  218. Chang-Yew Leow, C.; Gerondakis, S.; Spencer, A. MEK Inhibitors as a Chemotherapeutic Intervention in Multiple Myeloma. Blood Cancer J. 2013, 3, e105. [Google Scholar] [CrossRef]
  219. Steinbrunn, T.; Stühmer, T.; Sayehli, C.; Chatterjee, M.; Einsele, H.; Bargou, R.C. Combined Targeting of MEK/MAPK and PI3K/AKt Signalling in Multiple Myeloma. Br. J. Haematol. 2012, 159, 430–440. [Google Scholar] [CrossRef]
  220. Trudel, S.; Bahlis, N.J.; Venner, C.P.; Hay, A.E.; Kis, O.; Chow, S.; Li, Z.H.; Wei, E.N.; Wang, L.; Tran, C.; et al. Biomarker Driven Phase II Clinical Trial of Trametinib in Relapsed/Refractory Multiple Myeloma with Sequential Addition of the AKT Inhibitor, GSK2141795 at Time of Disease Progression to Overcome Treatment Failure: A Trial of the Princess Margaret Phase II Consortium. Blood 2016, 128, 4526. [Google Scholar] [CrossRef]
  221. Hideshima, T.; Akiyama, M.; Hayashi, T.; Richardson, P.; Schlossman, R.; Chauhan, D.; Anderson, K.C. Targeting P38 MAPK Inhibits Multiple Myeloma Cell Growth in the Bone Marrow Milieu. Blood 2003, 101, 703–705. [Google Scholar] [CrossRef]
  222. Reddy, M.; Ma, J.Y.; Kerr, I.; Henjarappa, N.; Navas, T.; Chakravarty, J.; Dugar, S.; Higgins, L.S.; Protter, A.A.; Medicherla, S. P38 Alpha MAP Kinase Inhibitor Inhibits Human Myeloma Cell Growth in Vivo. Blood 2004, 104, 4936. [Google Scholar] [CrossRef]
  223. Fu, J.; Li, S.; Liu, G.; Ma, H.; Marcireau, C.; Mapara, M.; Lentzsch, S. Synergistic Anti-Tumor Effects in RAS Mut Multiple Myeloma By Targeting MAP4K2 and RAS Pathways. Blood 2023, 142, 6618. [Google Scholar] [CrossRef]
  224. Malamos, P.; Papanikolaou, C.; Gavriatopoulou, M.; Dimopoulos, M.A.; Terpos, E.; Souliotis, V.L. The Interplay between the DNA Damage Response (DDR) Network and the Mitogen-Activated Protein Kinase (MAPK) Signaling Pathway in Multiple Myeloma. IJMS 2024, 25, 6991. [Google Scholar] [CrossRef] [PubMed]
  225. Alagpulinsa, D.A.; Ayyadevara, S.; Yaccoby, S.; Shmookler Reis, R.J. A Cyclin-Dependent Kinase Inhibitor, Dinaciclib, Impairs Homologous Recombination and Sensitizes Multiple Myeloma Cells to PARP Inhibition. Mol. Cancer Ther. 2016, 15, 241–250. [Google Scholar] [CrossRef]
  226. Botrugno, O.A.; Bianchessi, S.; Zambroni, D.; Frenquelli, M.; Belloni, D.; Bongiovanni, L.; Girlanda, S.; Di Terlizzi, S.; Ferrarini, M.; Ferrero, E.; et al. ATR Addiction in Multiple Myeloma: Synthetic Lethal Approaches Exploiting Established Therapies. haematologica 2019, 105, 2440–2447. [Google Scholar] [CrossRef]
  227. Xu, A.; Mazumder, A.; Borowiec, J.A. Cell Line-Dependent Synergy Between the PARP Inhibitor Veliparib and the Proteasome Inhibitor Bortezomib in the Killing of Myeloma Cells. Blood 2015, 126, 5556. [Google Scholar] [CrossRef]
  228. Bernasconi, E.; Gaudio, E.; Lejeune, P.; Tarantelli, C.; Cascione, L.; Kwee, I.; Spriano, F.; Rinaldi, A.; Mensah, A.A.; Chung, E.; et al. Preclinical Evaluation of the BET Bromodomain Inhibitor BAY 1238097 for the Treatment of Lymphoma. Br. J. Haematol. 2017, 178, 936–948. [Google Scholar] [CrossRef]
  229. Carew, J.S.; Espitia, C.M.; Zhao, W.; Visconte, V.; Kelly, K.R.; Nawrocki, S.T. Antagonizing HDAC6 Activity: A New Strategy to Maximize the Anti-Myeloma Effects of BET Inhibition. Blood 2015, 126, 3022. [Google Scholar] [CrossRef]
  230. Huang, Y.; Liu, N.; Pan, Z.; Li, Z.; Sheng, C. BET–HDAC Dual Inhibitors for Combinational Treatment of Breast Cancer and Concurrent Candidiasis. J. Med. Chem. 2023, 66, 1239–1253. [Google Scholar] [CrossRef]
  231. Bekri, S.; Rodney-Sandy, R.; Gruenstein, D.; Mei, A.; Bogen, B.; Castle, J.; Levey, D.; Cho, H.J. Neoantigen Vaccine-Induced CD4 T Cells Confer Protective Immunity in a Mouse Model of Multiple Myeloma through Activation of CD8 T Cells against Non-Vaccine, Tumor-Associated Antigens. J. Immunother. Cancer 2022, 10, e003572. [Google Scholar] [CrossRef]
  232. Kodysh, J.; Marron, T.; Rubinsteyn, A.; O’Donnell, T.; Finnigan, J.; Blazquez, A.; Saxena, M.; Meseck, M.; Friedlander, P.; Bhardwaj, N. Abstract CT173: PGV-001: A Phase I Trial of a Multipeptide Personalized Neoantigen Vaccine in the Adjuvant Setting. Cancer Res. 2020, 80, CT173. [Google Scholar] [CrossRef]
  233. Marron, T.U.; Saxena, M.; Bhardwaj, N.; Meseck, M.; Rubinsteyn, A.; Finnigan, J.; Kodysh, J.; Blazquez, A.; O’Donnel, T.; Galsky, M.; et al. Abstract LB048: An Adjuvant Personalized Neoantigen Peptide Vaccine for the Treatment of Malignancies (PGV-001). Cancer Res. 2021, 81, LB048. [Google Scholar] [CrossRef]
  234. Perumal, D.; Imai, N.; Laganà, A.; Finnigan, J.; Melnekoff, D.; Leshchenko, V.V.; Solovyov, A.; Madduri, D.; Chari, A.; Cho, H.J.; et al. Mutation-Derived Neoantigen-Specific T-Cell Responses in Multiple Myeloma. Clin. Cancer Res. 2020, 26, 450–464. [Google Scholar] [CrossRef] [PubMed]
  235. Jørgensen, N.G.; Ahmad, S.M.; Abildgaard, N.; Straten, P.T.; Svane, I.M.; Andersen, M.H.; Knudsen, L.M. Peptide Vaccination against Multiple Myeloma Using Peptides Derived from Anti-Apoptotic Proteins: A Phase I Trial. Stem Cell Investig. 2016, 3, 95. [Google Scholar] [CrossRef] [PubMed]
  236. Bae, J.; Smith, R.; Daley, J.; Mimura, N.; Tai, Y.-T.; Anderson, K.C.; Munshi, N.C. Myeloma-Specific Multiple Peptides Able to Generate Cytotoxic T Lymphocytes: A Potential Therapeutic Application in Multiple Myeloma and Other Plasma Cell Disorders. Clin. Cancer Res. 2012, 18, 4850–4860. [Google Scholar] [CrossRef]
  237. Bekri, S.; Uduman, M.; Gruenstein, D.; Mei, A.H.-C.; Tung, K.; Rodney-Sandy, R.; Bogen, B.; Buell, J.; Stein, R.; Doherty, K.; et al. Neoantigen Synthetic Peptide Vaccine for Multiple Myeloma Elicits T Cell Immunity in a Pre-Clinical Model. Blood 2017, 130, 1868. [Google Scholar]
  238. Bae, J.; Cowens, K.; Kimmelman, A.; Rowell, S.; Lazo, S.; Keskin, D.; Munshi, N.; Anderson, K. OP01-5 Combination Immunotherapy Targeting XBP1, CD138 and CS1 with Cancer Vaccine, Checkpoint Inhibitors, Immune Modulator, and Epigenetic Regulator to Treat Patients with Myeloma or Solid Tumours: Clinical Applications. ESMO Open 2023, 8, 102126. [Google Scholar] [CrossRef]
  239. Gorgun, G.T.; Whitehill, G.; Anderson, J.L.; Hideshima, T.; Laubach, J.P.; Raje, N.; Munshi, N.C.; Richardson, P.G.; Anderson, K.C. Myeloid Derived Suppressor Cells (MDSCs) Regulate Tumor Growth, Immune Response and Regulatory T Cell (Treg) Development in the Multiple Myeloma Bone Marrow Microenvironment. Blood 2012, 120, 565. [Google Scholar] [CrossRef]
  240. Malek, E.; De Lima, M.; Letterio, J.J.; Kim, B.-G.; Finke, J.H.; Driscoll, J.J.; Giralt, S.A. Myeloid-Derived Suppressor Cells: The Green Light for Myeloma Immune Escape. Blood Rev. 2016, 30, 341–348. [Google Scholar] [CrossRef]
  241. Topal Gorgun, G.; Ohguchi, H.; Hideshima, T.; Tai, Y.-T.; Raje, N.; Munshi, N.C.; Richardson, P.G.; Laubach, J.P.; Anderson, K.C. Inhibition Of Myeloid Derived Suppressor Cells (MDSC) In The Multiple Myeloma Bone Marrow Microenvironment. Blood 2013, 122, 3089. [Google Scholar] [CrossRef]
  242. Botta, C.; Gullà, A.; Correale, P.; Tagliaferri, P.; Tassone, P. Myeloid-Derived Suppressor Cells in Multiple Myeloma: Pre-Clinical Research and Translational Opportunities. Front. Oncol. 2014, 4, 348. [Google Scholar] [CrossRef]
  243. Albu, D.I.; Wang, Z.; Huang, K.-C.; Wu, J.; Twine, N.; Leacu, S.; Ingersoll, C.; Parent, L.; Lee, W.; Liu, D.; et al. EP4 Antagonism by E7046 Diminishes Myeloid Immunosuppression and Synergizes with Treg-Reducing IL-2-Diphtheria Toxin Fusion Protein in Restoring Anti-Tumor Immunity. OncoImmunology 2017, 6, e1338239. [Google Scholar] [CrossRef]
  244. Giannotta, C.; Autino, F.; Massaia, M. The Immune Suppressive Tumor Microenvironment in Multiple Myeloma: The Contribution of Myeloid-Derived Suppressor Cells. Front. Immunol. 2023, 13, 1102471. [Google Scholar] [CrossRef]
  245. Park, J.J.H.; Hsu, G.; Siden, E.G.; Thorlund, K.; Mills, E.J. An Overview of Precision Oncology Basket and Umbrella Trials for Clinicians. CA Cancer J. Clin. 2020, 70, 125–137. [Google Scholar] [CrossRef]
  246. Kesari, H.V.; Ravi, R. Novel Study Designs in Precision Medicine—Basket, Umbrella and PlatformTrials. CRCEP 2022, 17, 114–121. [Google Scholar] [CrossRef]
  247. González-Calle, V.; Keane, N.; Braggio, E.; Fonseca, R. Precision Medicine in Myeloma: Challenges in Defining an Actionable Approach. Clin. Lymphoma Myeloma Leuk. 2017, 17, 621–630. [Google Scholar] [CrossRef] [PubMed]
  248. Moore, D.C.; Guinigundo, A.S. Biomarker-Driven Oncology Clinical Trials: Novel Designs in the Era of Precision Medicine. JADPRO 2023, 14, 9–13. [Google Scholar] [CrossRef] [PubMed]
  249. Pan, D.; Richter, J. Where We Stand with Precision Therapeutics in Myeloma: Prosperity, Promises, and Pipedreams. Front. Oncol. 2022, 11, 819127. [Google Scholar] [CrossRef]
  250. Zweegman, S.; Engelhardt, M.; Larocca, A. Elderly Patients with Multiple Myeloma: Towards a Frailty Approach? Curr. Opin. Oncol. 2017, 29, 315–321. [Google Scholar] [CrossRef]
  251. Sutanto, H. Tackling Polypharmacy in Geriatric Patients: Is Increasing Physicians’ Awareness Adequate? Arch. Gerontol. Geriatr. Plus 2025, 2, 100185. [Google Scholar] [CrossRef]
  252. Möller, M.-D.; Gengenbach, L.; Graziani, G.; Greil, C.; Wäsch, R.; Engelhardt, M. Geriatric Assessments and Frailty Scores in Multiple Myeloma Patients: A Needed Tool for Individualized Treatment? Curr. Opin. Oncol. 2021, 33, 648–657. [Google Scholar] [CrossRef]
  253. Larocca, A.; Palumbo, A. Frail Patients with Newly Diagnosed Multiple Myeloma. In Neoplastic Diseases of the Blood; Wiernik, P.H., Dutcher, J.P., Gertz, M.A., Eds.; Springer International Publishing: Cham, Switzerland, 2018; pp. 539–549. ISBN 978-3-319-64263-5. [Google Scholar]
  254. Bao, L.; Wang, Y.-T.; Liu, P.; Lu, M.-Q.; Zhuang, J.-L.; Zhang, M.; Xia, Z.-J.; Li, Z.-L.; Yang, Y.; Yan, Z.-Y.; et al. Ixazomib-Based Frontline Therapy Followed by Ixazomib Maintenance in Frail Elderly Newly Diagnosed with Multiple Myeloma: A Prospective Multicenter Study. EClinicalMedicine 2024, 68, 102431. [Google Scholar] [CrossRef]
  255. Salazar, A.S.; Recinos, L.M.; Mian, H.S.; Stoll, C.; Simon, L.E.; Sekhon, S.; Colditz, G.A.; Wildes, T.M. Geriatric Assessment and Frailty Scores Predict Mortality in Myeloma: Systematic Review and Meta-Analysis. Clin. Lymphoma Myeloma Leuk. 2019, 19, 488–496.e6. [Google Scholar] [CrossRef]
  256. Smits, F.; Nasserinejad, K.; Levin, M.-D.; Timmers, G.J.; De Waal, E.G.M.; Regelink, J.C.; Velders, G.A.; Westerman, M.; Durdu-Rayman, N.; Last-Koopmans, S.; et al. Dynamics of Geriatric Impairments Are Associated with Risk of Treatment Discontinuation in Elderly Patients with Multiple Myeloma. Blood 2024, 144, 3330. [Google Scholar] [CrossRef]
  257. Nathwani, N.; Kurtin, S.E.; Lipe, B.; Mohile, S.G.; Catamero, D.D.; Wujcik, D.; Birchard, K.; Davis, A.; Dudley, W.; Stricker, C.T.; et al. Integrating Touchscreen-Based Geriatric Assessment and Frailty Screening for Adults with Multiple Myeloma to Drive Personalized Treatment Decisions. JCO Oncol. Pract. 2020, 16, e92–e99. [Google Scholar] [CrossRef]
  258. Greenberg, A.J.; Vachon, C.M.; Rajkumar, S.V. Disparities in the Prevalence, Pathogenesis and Progression of Monoclonal Gammopathy of Undetermined Significance and Multiple Myeloma between Blacks and Whites. Leukemia 2012, 26, 609–614. [Google Scholar] [CrossRef]
  259. Ailawadhi, S.; Jacobus, S.; Sexton, R.; Stewart, A.K.; Dispenzieri, A.; Hussein, M.A.; Zonder, J.A.; Crowley, J.; Hoering, A.; Barlogie, B.; et al. Disease and Outcome Disparities in Multiple Myeloma: Exploring the Role of Race/Ethnicity in the Cooperative Group Clinical Trials. Blood Cancer J. 2018, 8, 67. [Google Scholar] [CrossRef]
  260. Kazandjian, D.; Hill, E.; Hultcrantz, M.; Rustad, E.H.; Yellapantula, V.; Akhlaghi, T.; Korde, N.; Mailankody, S.; Dew, A.; Papaemmanuil, E.; et al. Molecular Underpinnings of Clinical Disparity Patterns in African American vs. Caucasian American Multiple Myeloma Patients. Blood Cancer J. 2019, 9, 15. [Google Scholar] [CrossRef] [PubMed]
  261. Dumont, E.L.; Han, L.; Kodali, S.; Aptekmann, A.; Carter-Bawa, L.; Feinman, R.; Tycko, B.; Siegel, D.S.; Goy, A.; Kaplan, P.; et al. Genetic Ethnicity and Hypertension Epistatic Interaction Underlying Racial Disparities in US Multiple Myeloma Susceptibility. medRxiv 2024. [Google Scholar] [CrossRef]
  262. Pierre, A.; Williams, T. African American Patients with Multiple Myeloma: Optimizing Care to Decrease Racial Disparities. CJON 2020, 24, 439–443. [Google Scholar] [CrossRef]
  263. Guadamuz, J.S.; Rohrer, R.; Mohyuddin, G.R.; Chiu, B.C.-H.; Patel, P.R.; Sweiss, K.; Seymour, E.; Sborov, D.; Calip, G.S. Abstract 3669: Racial/Ethnic Disparities in Treatment with Bone-Modifying Agents among Newly Diagnosed Multiple Myeloma Patients. Cancer Res. 2022, 82, 3669. [Google Scholar] [CrossRef]
  264. Jasaraj, R.; Gaire, S.; Proskuriakova, E.; Jasaraj, R.; Keruakous, A.R. Racial Bias in Multiple Myeloma Clinical Trials: Racial Minorities Continue to Miss out on Treatment Advances. JCO 2023, 41, e20009. [Google Scholar] [CrossRef]
  265. Kelly, K.R.; Yang, D.; Sharma, M.; Roy, V.; Chanan-Khan, A.A.; Ailawadhi, S. Persistent Racial/Ethnic Disparities in Outcomes for Multiple Myeloma: A SEER-Database Update. Blood 2016, 128, 1191. [Google Scholar] [CrossRef]
  266. Jain, T.; Fonseca, R.; Chen, R.; Patel, R.; Jani, P.; De La Calle, V.G.; Meghji, Z.; Hoffman, J.E.; Alencar, A.J.; Kelly, K.R.; et al. Racial Differences in Disease Characteristics: Understanding Multiple Myeloma in Hispanics. Blood 2017, 130, 864. [Google Scholar] [CrossRef]
  267. Moreau, P.; Kumar, S.K.; San Miguel, J.; Davies, F.; Zamagni, E.; Bahlis, N.; Ludwig, H.; Mikhael, J.; Terpos, E.; Schjesvold, F.; et al. Treatment of Relapsed and Refractory Multiple Myeloma: Recommendations from the International Myeloma Working Group. Lancet Oncol. 2021, 22, e105–e118. [Google Scholar] [CrossRef]
  268. Miguel, J.F.S.; Mateos, M.-V. Current Challenges in the Management of Patients with Relapsed/Refractory Multiple Myeloma. Oncology 2011, 25 (Suppl. S2), 10–18. [Google Scholar]
  269. Offidani, M.; Corvatta, L.; Morabito, F.; Gentile, M.; Musto, P.; Leoni, P.; Palumbo, A. How to Treat Patients with Relapsed/Refractory Multiple Myeloma: Evidence-Based Information and Opinions. Expert Opin. Investig. Drugs 2011, 20, 779–793. [Google Scholar] [CrossRef]
  270. Ni, B.; Hou, J. Promising Therapeutic Approaches for Relapsed/Refractory Multiple Myeloma. Hematology 2022, 27, 343–352. [Google Scholar] [CrossRef]
  271. Richardson, P.G.; Laubach, J.P.; Schlossman, R.L.; Ghobrial, I.M.; Redman, K.C.; Mckenney, M.; Warren, D.; Noonan, K.; Lunde, L.; Doss, D.; et al. The Potential Benefits of Participating in Early-phase Clinical Trials in Multiple Myeloma: Long-term Remission in a Patient with Relapsed Multiple Myeloma Treated with 90 Cycles of Lenalidomide and Bortezomib. Eur. J. Haematol. 2012, 88, 446–449. [Google Scholar] [CrossRef]
  272. Clark, C.A.; Cornell, R.F.; Scott, E.C.; Chung, J.; Costa, L.J. Management of Relapsed and Refractory Multiple Myeloma in Modern Times: Incorporating New Agents into Decision-making. Am. J. Hematol. 2016, 91, 1044–1051. [Google Scholar] [CrossRef]
Figure 1. Clonal evolution and disease progression in the multiple myeloma continuum. The biologic continuum of MM development starts from a post-germinal center (Post-GC) B cell through MGUS, SMM, overt MM, and ultimately extramedullary disease (EMD). Key initiating events include hyperdiploidy and IgH translocations. Progression from MGUS to MM is driven by mutations such as KRAS, MYC activation, TP53 inactivation, and increased genomic instability. Further disease progression involves secondary genetic alterations, notably TP53 mutations and deletions such as Del(17p). Clonal evolution is driven by intrinsic genetic instability, clonal selection, adaptive mutations, and external selective pressures including the immune system, tumor microenvironment, and therapeutic interventions.
Figure 1. Clonal evolution and disease progression in the multiple myeloma continuum. The biologic continuum of MM development starts from a post-germinal center (Post-GC) B cell through MGUS, SMM, overt MM, and ultimately extramedullary disease (EMD). Key initiating events include hyperdiploidy and IgH translocations. Progression from MGUS to MM is driven by mutations such as KRAS, MYC activation, TP53 inactivation, and increased genomic instability. Further disease progression involves secondary genetic alterations, notably TP53 mutations and deletions such as Del(17p). Clonal evolution is driven by intrinsic genetic instability, clonal selection, adaptive mutations, and external selective pressures including the immune system, tumor microenvironment, and therapeutic interventions.
Hemato 06 00029 g001
Figure 2. Hallmarks of malignant plasma cell transformation from MGUS to MM. Uncontrolled proliferation reflects the autonomous expansion of malignant plasma cells. Evasion of apoptosis enables these cells to resist programmed cell death. Genomic instability accelerates mutation accumulation and clonal evolution. Bone marrow microenvironment dependency highlights the supportive role of stromal and immune cell interactions in sustaining myeloma growth. Immune evasion allows malignant cells to avoid detection and elimination by the immune system. Enhanced angiogenesis supplies nutrients and oxygen through neovascularization. Metabolic reprogramming supports the increased energy demands of tumor cells. Epigenetic dysregulation contributes to abnormal gene expression without altering the DNA sequence. Aberrant cellular adhesion and migration facilitate disease dissemination and bone marrow colonization. Sustained identity with aberrant differentiation reflects the retention of plasma cell features alongside malignant transformation.
Figure 2. Hallmarks of malignant plasma cell transformation from MGUS to MM. Uncontrolled proliferation reflects the autonomous expansion of malignant plasma cells. Evasion of apoptosis enables these cells to resist programmed cell death. Genomic instability accelerates mutation accumulation and clonal evolution. Bone marrow microenvironment dependency highlights the supportive role of stromal and immune cell interactions in sustaining myeloma growth. Immune evasion allows malignant cells to avoid detection and elimination by the immune system. Enhanced angiogenesis supplies nutrients and oxygen through neovascularization. Metabolic reprogramming supports the increased energy demands of tumor cells. Epigenetic dysregulation contributes to abnormal gene expression without altering the DNA sequence. Aberrant cellular adhesion and migration facilitate disease dissemination and bone marrow colonization. Sustained identity with aberrant differentiation reflects the retention of plasma cell features alongside malignant transformation.
Hemato 06 00029 g002
Figure 3. Immunotherapeutic targets and strategies in multiple myeloma. Monoclonal antibodies such as daratumumab and isatuximab target CD38, while elotuzumab targets SLAMF7, promoting immune-mediated cytotoxicity. Immune checkpoint inhibitors (anti-PD-L1/PD-1) aim to restore T-cell function by disrupting the PD-1/PD-L1 axis. CAR-T-cell therapies directed against CD19 and BCMA enhance cytotoxic T-cell responses. Bispecific T-cell engagers (BiTEs), including AMG420 and blinatumomab, bridge T cells (via CD3) to MM cells expressing BCMA. Immunomodulatory drugs (thalidomide, lenalidomide, pomalidomide) act by altering the bone marrow stromal cell (BMSC) interaction and enhancing immune surveillance. Multipeptide vaccines target tumor-associated antigens (MAGE, WT-1, XBP1) presented by dendritic cells, aiming to prime anti-myeloma immune responses.
Figure 3. Immunotherapeutic targets and strategies in multiple myeloma. Monoclonal antibodies such as daratumumab and isatuximab target CD38, while elotuzumab targets SLAMF7, promoting immune-mediated cytotoxicity. Immune checkpoint inhibitors (anti-PD-L1/PD-1) aim to restore T-cell function by disrupting the PD-1/PD-L1 axis. CAR-T-cell therapies directed against CD19 and BCMA enhance cytotoxic T-cell responses. Bispecific T-cell engagers (BiTEs), including AMG420 and blinatumomab, bridge T cells (via CD3) to MM cells expressing BCMA. Immunomodulatory drugs (thalidomide, lenalidomide, pomalidomide) act by altering the bone marrow stromal cell (BMSC) interaction and enhancing immune surveillance. Multipeptide vaccines target tumor-associated antigens (MAGE, WT-1, XBP1) presented by dendritic cells, aiming to prime anti-myeloma immune responses.
Hemato 06 00029 g003
Figure 4. Comparison of basket and umbrella trial designs in precision oncology. In a basket trial (left panel), patients with various tumor types but a shared molecular alteration are enrolled to evaluate the efficacy of a targeted therapy independent of cancer histology. This approach enables assessment of treatment effects based on genomic or molecular characteristics rather than tumor origin. In contrast, an umbrella trial (right panel) focuses on a single tumor type—such as multiple myeloma (MM)—and stratifies patients based on distinct molecular subtypes or biomarkers. Each subgroup receives a different, personalized therapy (e.g., Therapy A, B, or C), enabling simultaneous evaluation of multiple targeted treatments within a single cancer type. These trial designs optimize therapeutic matching and accelerate the development of individualized cancer therapies.
Figure 4. Comparison of basket and umbrella trial designs in precision oncology. In a basket trial (left panel), patients with various tumor types but a shared molecular alteration are enrolled to evaluate the efficacy of a targeted therapy independent of cancer histology. This approach enables assessment of treatment effects based on genomic or molecular characteristics rather than tumor origin. In contrast, an umbrella trial (right panel) focuses on a single tumor type—such as multiple myeloma (MM)—and stratifies patients based on distinct molecular subtypes or biomarkers. Each subgroup receives a different, personalized therapy (e.g., Therapy A, B, or C), enabling simultaneous evaluation of multiple targeted treatments within a single cancer type. These trial designs optimize therapeutic matching and accelerate the development of individualized cancer therapies.
Hemato 06 00029 g004
Table 1. Comparison of linear vs. branching clonal evolution in multiple myeloma.
Table 1. Comparison of linear vs. branching clonal evolution in multiple myeloma.
FeatureLinear EvolutionBranching Evolution
DefinitionStepwise accumulation of mutations in a single dominant cloneMultiple subclones diverge from a common ancestor and evolve in parallel
Clonal dynamicsOne clone sequentially replaces the previous one (clonal sweeps)Several clones coexist and compete; dominance can shift over time
Genetic trajectoryUnidirectional with sequential driver mutationsDiverse mutational paths; subclones may acquire different mutations independently
Temporal patternProgressive and orderedNon-linear and dynamic
Driver mutation DistributionAccumulated in a linear fashion within a single lineageFound in distinct subclones; can appear independently
Influence of therapySelective pressure can accelerate emergence of the next dominant cloneSelective pressure alters subclone proportions; resistant clones may expand
Clinical presentationOften associated with more aggressive disease progression and shorter survivalMore heterogeneous presentation; some subclones may respond differently to therapy
Frequency in MMLess common (~15–25%)More common (~60–70%)
DetectionEasier to detect using bulk sequencing due to dominant cloneRequires single-cell sequencing or longitudinal sampling to fully resolve
Example markersTP53 inactivation, high mutational loadDivergent RAS mutations in separate subclones
Prognostic implicationOften associated with poor prognosis and rapid relapseVariable prognosis; can enable early detection of resistance pathways
Therapeutic implicationsMay benefit from aggressive upfront treatment strategiesSupports adaptive or combination therapy approaches to target multiple clones
Table 2. Components of the bone marrow microenvironment involved in multiple myeloma.
Table 2. Components of the bone marrow microenvironment involved in multiple myeloma.
ComponentCell Type/MoleculeRole in MM PathogenesisKey Molecular Mechanisms/Pathways
Stromal cellsBone marrow stromal cells (BMSCs)Promote MM cell survival, proliferation, and drug resistanceIL-6, SDF-1/CXCL12 → CXCR4, adhesion molecules (VCAM-1, VLA-4), NF-κB activation
OsteoclastsMultinucleated bone-resorbing cellsPromote bone destruction and release growth factors for MM cellsRANK/RANKL, M-CSF, IL-6, TGF-β; release of IGF-1 and calcium enhances MM growth
OsteoblastsBone-forming cellsSuppressed in MM, leading to bone lesionsInhibition via DKK1, sFRP2, and activin A; Wnt pathway inhibition
Endothelial cellsVascular endothelial cellsPromote angiogenesis and MM progressionVEGF, FGF-2, HIF-1α pathways; MM cells stimulate angiogenic switch
Immune cells—Myeloid lineageMyeloid-derived suppressor cells (MDSCs)Suppress anti-MM immunity, support MM cell survivalArginase-1, ROS, IL-10, TGF-β; inhibit cytotoxic T cells and NK cells
Immune cells—Lymphoid lineageRegulatory T cells (Tregs)Inhibit effective immune responses against MMIL-10, TGF-β, CTLA-4; suppression of CD8+ T-cell activation
Dendritic cellsAntigen-presenting cellsImpaired antigen presentation; promote immune tolerance in MMDownregulation of costimulatory molecules, IL-6, IDO expression
MacrophagesTumor-associated macrophages (TAMs)Support MM growth and immune evasionIL-6, TNF-α, CD163 expression; M2 polarization promotes tumor-supportive environment
Mesenchymal stem cells (MSCs)Multipotent stromal progenitorsGive rise to BMSCs; modulate inflammation and MM nicheSecretion of IL-6, CXCL12; direct interactions with MM cells via adhesion molecules
AdipocytesBone marrow fat cellsProvide metabolic support and cytokines for MM cellsLeptin, adiponectin, IL-6, lipolysis-derived fatty acids; affect metabolism and resistance
Extracellular matrix (ECM)Fibronectin, collagen, laminin, hyaluronanPromotes MM adhesion, survival, and drug resistanceIntegrins (VLA-4/VCAM-1), FAK, PI3K/AKT, ERK signaling; CAM-DR (cell adhesion-mediated drug resistance)
Cytokines and chemokinesIL-6, IL-10, TGF-β, SDF-1 (CXCL12), TNF-αMediate pro-survival, pro-migratory, and immunosuppressive signalsActivation of JAK/STAT3, MAPK, NF-κB, and PI3K/AKT pathways
Growth factorsVEGF, IGF-1, BAFF, APRIL, FGFEnhance angiogenesis, MM cell proliferation, and immune evasionBinding to receptors on MM cells (e.g., IGF-1R, VEGFR), MAPK/ERK and AKT activation
Exosomes/extracellular vesiclesSecreted by MM cells and stromal cellsMediate intercellular communication, resistance, and immune modulationTransfer of miRNAs, proteins; enhance angiogenesis, suppress T-cell function
Table 3. Comparison of hyperdiploid vs. non-hyperdiploid multiple myeloma.
Table 3. Comparison of hyperdiploid vs. non-hyperdiploid multiple myeloma.
FeatureHyperdiploid Multiple Myeloma (H-MM)Non-Hyperdiploid Multiple Myeloma (NH-MM)
DefinitionPresence of trisomies (usually odd-numbered chromosomes)Characterized by structural abnormalities, especially IgH translocations
Common cytogenetic featuresTrisomies of chromosomes 3, 5, 7, 9, 11, 15, 19, 21IgH translocations: t(11;14), t(4;14), t(14;16), t(14;20), t(6;14)
Frequency~50–60% of newly diagnosed MM~30–40% of newly diagnosed MM
Oncogenic mechanismGene dosage effect from trisomiesOncogene activation via enhancer hijacking (e.g., CCND1, FGFR3/MMSET, MAF)
Gene expression profileEnrichment in genes linked to protein synthesis and oxidative phosphorylationEnrichment in cell cycle, proliferation, and oncogenic transcription factors
MYC deregulationLess frequentMore frequent, especially in t(4;14), t(14;16) subtypes
Cyclin D deregulationVia gene dosage (CCND1, CCND2)Via IgH translocations (e.g., t(11;14) → CCND1 overexpression)
PrognosisGenerally more favorableVariable, often poorer (especially with t(4;14), del(17p), or 1q gain)
Associated risk abnormalitiesFewer high-risk featuresMore frequent del(17p), 1q gain, and complex karyotypes
Response to therapyGood response to standard therapyOften requires risk-adapted or novel agents (e.g., bortezomib in t(4;14))
Microenvironmental interactionLess dependent on niche signalingGreater dependence on IL-6 and stromal support in early clones
Progression patternTypically slower progression from MGUSCan have rapid evolution from SMM or de novo aggressive disease
Common clinical featuresOlder age, indolent courseYounger patients more common; higher tumor burden and extramedullary disease
Table 4. IgH translocations in multiple myeloma.
Table 4. IgH translocations in multiple myeloma.
TranslocationPartner Gene/LocusFrequencyMolecular ConsequenceAssociated Clinical Features/Risk
t(11;14)(q13;q32)CCND1 (cyclin D1)~15–20%Overexpression of cyclin D1, promoting G1/S cell cycle entryStandard-risk; enriched in light chain amyloidosis, lymphoplasmacytic morphology, and BCL-2 dependency (venetoclax-responsive)
t(4;14)(p16;q32)FGFR3/MMSET (NSD2)~15%Dual activation: FGFR3 (MAPK signaling) and MMSET (epigenetic modulation)High-risk; associated with poor prognosis, early relapse, and benefit from bortezomib-based therapy
t(14;16)(q32;q23)MAF (v-maf musculoaponeurotic fibrosarcoma)~5–7%Upregulation of MAF transcription factor, altering adhesion and migration genesHigh-risk; associated with extramedullary disease and plasma cell leukemia-like features
t(14;20)(q32;q12)MAFB<2%Overexpression of MAFB, disrupting plasma cell differentiationHigh-risk; rare but aggressive, with poor response to conventional therapy
t(6;14)(p21;q32)CCND3 (cyclin D3)<1–2%Overexpression of cyclin D3, promoting cell cycle progressionRare; clinical significance not fully established; intermediate prognosis
t(8;14)(q24;q32)MYC (secondary event)~15–30% (progression)Enhancer hijacking leads to MYC overexpressionSecondary translocation in disease progression; associated with high tumor burden and relapse
t(9;14)(p13;q32)PAX5Very rareDisrupts B-cell differentiation pathwayExtremely rare in MM; more common in lymphomas
Table 5. Comparison of RAS pathway mutations in multiple myeloma.
Table 5. Comparison of RAS pathway mutations in multiple myeloma.
FeatureKRAS MutationNRAS MutationBRAF Mutation
Gene locationChromosome 12p12.1Chromosome 1p13.2Chromosome 7q34
Frequency in MM~20–25%~20–25%~4–8%
Common mutation hotspotsG12D, G12V, G13D, Q61HQ61K, Q61R, G12D, G13RV600E (most common), K601N
Activation effectConstitutive activation of MAPK and PI3K/AKTPreferential activation of MAPK signalingStrong MAPK pathway activation via MEK/ERK
Pathway preferenceBoth MAPK and PI3K pathwaysMAPK (ERK-driven proliferation)MAPK pathway exclusively
Mutual exclusivityGenerally mutually exclusive with NRAS/BRAFRare co-occurrence with KRAS/BRAFUsually exclusive of RAS mutations
Prognostic impactVariable; may be associated with relapseMay be associated with early progressionV600E associated with poor prognosis and high-risk disease
Clonal distributionOften subclonal at diagnosis, may become clonal at relapseOften clonal and persistent through disease courseCan be early or late event
Functional role in MMPromotes cell proliferation, resistance to apoptosisDrives proliferation and survivalPromotes rapid proliferation and metabolic changes
Response to MEK inhibitorsVariable; better response in dual PI3K/MEK inhibitionSimilar to KRAS; requires combination therapyMore responsive to BRAF/MEK inhibitors (e.g., vemurafenib + cobimetinib)
Preclinical targetabilityRequires combined targeting of PI3K and MEKLimited success with MEK inhibitors aloneV600E mutation shows sensitivity to BRAF inhibitors in vitro
Clinical trial implicationsEnriched in biomarker-driven MEK/AKT inhibitor studiesEvaluated in MAPK pathway-targeted trialsEnrolled in BRAF inhibitor trials [64] (e.g., vemurafenib [VEM] + cobimetinib [COBI])
Table 6. IMWG diagnostic criteria for MGUS, SMM, and MM.
Table 6. IMWG diagnostic criteria for MGUS, SMM, and MM.
FeatureMGUSSMMMM
Clonal bone marrow plasma cells (BMPCs)<10%≥10% to <60%≥10% or biopsy-proven plasmacytoma
M-protein in serum (if present)<3 g/dL≥3 g/dL (if present) or urinary M-protein ≥500 mg/24 hAny level (often >3 g/dL)
Myeloma-defining EventsAbsentAbsentPresent (≥1 CRAB or SLiM criteria, see below)
CRAB features (myeloma-defining)NoneNoneAt least one of the following:
- C: hypercalcemia (Ca >11 mg/dL or >1 mg/dL above normal)
- R: renal insufficiency (Cr > 2 mg/dL or CrCl < 40 mL/min)
- A: anemia (Hb < 10 g/dL or >2 g/dL below normal)
- B: bone lesions (≥1 osteolytic lesion on imaging)
SLiM biomarkers (added in 2014)Not applicableNot applicablePresence of any of the following (even without CRAB):
- S: ≥60% clonal BMPCs
- Li: involved/uninvolved serum free light chain (FLC) ratio ≥100
- M: >1 focal lesion ≥5 mm on MRI
End-organ damageAbsentAbsentPresent or imminent (CRAB or SLiM features)
Progression risk (2-year)~1%~10% (standard SMM) to ~50% (high-risk SMM)100% (active disease requiring treatment)
Recommended managementObservation; no treatmentObservation or clinical trial (if high-risk)Initiate systemic therapy
Table 7. Comparison of ISS and R-ISS staging systems (including NGS and FISH integration).
Table 7. Comparison of ISS and R-ISS staging systems (including NGS and FISH integration).
FeatureInternational Staging System (ISS)Revised International Staging System (R-ISS)R2-ISS/NGS-Enhanced or FISH-Enhanced Systems [110]
Year Introduced20052015Emerging (post-2020), based on genomic insights
Primary PurposeEstimate prognosis using basic lab valuesRefine prognosis by integrating cytogenetics and LDHImprove risk stratification using NGS and/or advanced FISH data
Parameters Used- Serum β2-microglobulin
- Serum albumin
- ISS components
- Chromosomal abnormalities (via FISH)
- LDH level
- R-ISS components
- TP53 mutation status
- 1q gain, del(1p), RAS/BRAF mutations, others
Genetic/cytogenetic inputNoneYes (FISH): del(17p), t(4;14), t(14;16)Yes (NGS/FISH): includes extended high-risk markers (e.g., 1q gain, biallelic TP53 loss)
LDH incorporationNoYesYes
Stages defined- Stage I: β2M < 3.5 mg/L and serum albumin ≥ 3.5 g/dL
- Stage II: β2M < 3.5 mg/L; serum albumin < 3.5 g/dL; or β2M 3.5 to 5.5 mg/L, irrespective of serum albumin
- Stage III: β2M ≥ 5.5 mg/L
- Stage I: ISS I + no high-risk FISH + normal LDH
- Stage II: not R-ISS stage I or III
- Stage III: ISS III + high-risk FISH and/or high LDH
- More granular stratification (e.g., 4 groups in R2-ISS)
- Assigns scores to each risk marker for cumulative risk
Risk categories3 tiers (I, II, III)3 tiers (I, II, III)Often 4 tiers (e.g., R2-ISS groups I–IV)
Prognostic discriminationModerateImproved over ISSSuperior; captures genetic heterogeneity and clonal complexity
Clinical usefulnessStill used for baseline stagingWidely adopted in clinical trials and risk stratificationEmerging in precision medicine; useful in high-risk MM trials
LimitationsLacks genetic insightDoes not account for 1q gain, TP53 mutation, or multiple abnormalitiesRequires access to NGS and extended FISH; not yet fully standardized
Examples of added markersN/AFISH only: del(17p), t(4;14), t(14;16)NGS/FISH: TP53 mutations, biallelic inactivation, gain(1q21), del(1p), RAS/BRAF mutations, complex karyotype
Progression-free survival (PFS) and OS predictionLimited precisionBetter at stratifying high-risk patientsMost accurate; enables personalized risk-adapted therapy
Table 8. Comparison of MRD detection techniques in multiple myeloma.
Table 8. Comparison of MRD detection techniques in multiple myeloma.
FeatureNext-Generation Sequencing (NGS)Next-Generation Flow Cytometry (NGF)Allele-Specific Oligonucleotide PCR (ASO-PCR)Imaging (PET/CT, MRI)
Technology basisDNA sequencing of unique IgH rearrangementsMultiparameter flow cytometry with >8–10 color panelsPCR amplification of patient-specific IgH gene rearrangementFunctional/metabolic (PET/CT) and anatomical (MRI) imaging
Sample typeBone marrow (BM) aspirateBone marrow aspirateBone marrow aspirateWhole-body (bone marrow and extramedullary sites)
SensitivityUp to 10−6 (1 cell in 1 million)Up to 10−5 to 10−6~10−5~10−2 to 10−3 (lower sensitivity)
StandardizationHighly standardized (e.g., Adaptive ClonoSEQ®)Standardized by EuroFlow consortiumRequires individualized primers (complex setup)Variable depending on protocol and scanner
Turnaround timeModerate (few days to 1–2 weeks)Fast (24–48 h)Slow (requires pretreatment sample)Immediate (PET/CT ~same day; MRI may take longer)
Clonality requirementRequires identification of clone at diagnosisNot neededRequires diagnostic sample for primer designNot applicable
ApplicabilityApplicable in ≥90% of patientsApplicable in virtually all patientsApplicable in ~60–70%Applicable in all patients
Cost and accessibilityExpensive; limited to specialized labsModerate cost; increasingly available in tertiary centersCost-effective but labor-intensiveExpensive; PET/CT requires radiotracers
AdvantagesHighest sensitivity; highly quantitative; standardizedReal-time detection; rapid; does not require diagnostic sampleHighly sensitive; cost-effective for known cloneDetects extramedullary disease and focal lesions missed by BM sampling
LimitationsRequires diagnostic sample; BM only; not real-timeLimited to BM; may miss patchy or extramedullary diseaseTechnically demanding; not standardized; limited utilityLow sensitivity for minimal disease; can yield false positives due to inflammation
Detects extramedullary diseaseNoNoNoYes
Role in clinical trialsUsed in multiple trials as MRD endpoint (e.g., FORTE, GRIFFIN)Widely used in European MRD studiesRarely used in modern trialsUsed for imaging-based response and disease mapping
Regulatory approvalFDA-cleared (ClonoSEQ for MRD in MM)CE-marked (EuroFlow)Not FDA-approved for MRDStandard clinical use for imaging; not specific for MRD
Integration with IMWG MRD criteriaYesYesPartially (less common)Yes (for imaging-defined MRD negativity)
Table 9. Comparison of PET/CT, MRI, and WBLDCT in multiple myeloma imaging.
Table 9. Comparison of PET/CT, MRI, and WBLDCT in multiple myeloma imaging.
FeaturePET/CT (FDG-PET/CT)MRI (Whole-Body or Spine/Pelvis)WBLDCT (Whole-Body Low-Dose CT)
Imaging modalityFunctional (metabolic) + anatomical imagingAnatomical + functional (tissue contrast, edema detection)Anatomical (bone structure) only
Radiotracer or contrast18F-FDG (fluorodeoxyglucose)None required (but gadolinium can be used)No contrast or radiotracer required
Radiation exposureModerate (higher than WBLDCT)NoneLow-dose radiation
Sensitivity for bone lesionsModerate to high (especially for active lesions)Highest for diffuse and focal marrow infiltrationHigh for osteolytic lesions ≥5 mm
Sensitivity for early diseaseLess sensitive for non-metabolically active lesionsBest modality for early bone marrow involvementPoor at detecting marrow infiltration before bone destruction
Detection of extramedullary diseaseExcellentLimited to soft tissue near bonesNot reliable
Detection of focal lesionsYes (when metabolically active)Yes (even before bone destruction)Yes (if lytic and ≥5 mm)
Detection of diffuse infiltrationLimitedYes (via signal intensity on STIR/DWI sequences)No
Detection of healing or inactive lesionsNo (FDG-negative)Yes (but may not distinguish active from inactive)Yes (can see healed lesions, but not activity)
Use in MRD evaluationYes (for imaging-defined MRD negativity)Yes (e.g., ≥1 focal lesion ≥5 mm = SLiM criterion)No (not part of IMWG MRD criteria)
Turnaround timeModerate (requires radiotracer synthesis and uptake time)Fast (30–60 min; more with contrast)Fast (10–15 min)
AvailabilityModerate (depends on PET facility)Widely available in tertiary centersWidely available
LimitationsFalse negatives in low-FDG-avid MM; false positives from inflammationExpensive; time-consuming; contraindicated in severe renal failure (if contrast used)Can miss early or non-lytic disease
AdvantagesIdentifies active lesions and extramedullary diseaseBest for early marrow disease, ideal for SLiM criteriaBest for osteolytic lesion detection; quick and accessible
Table 10. Frontline pharmacotherapy for newly diagnosed multiple myeloma.
Table 10. Frontline pharmacotherapy for newly diagnosed multiple myeloma.
RegimenComponentsPatient PopulationMechanism of ActionKey Clinical TrialsResponse Rates/PFSSpecial Considerations
VRdBortezomib (V) + lenalidomide (R) + dexamethasone (d)Transplant-eligible and -ineligiblePI + IMiD + steroidSWOG S0777, IFM 2009ORR ~90%, CR ~30–40%, median PFS ~43–50 monthsStandard of care; neuropathy with bortezomib SC > IV preferred
Dara-VRdDaratumumab (Dara) + VRdTransplant-eligible (per FDA/EMA)Anti-CD38 mAb + PI + IMiD + steroidGRIFFIN, PERSEUSsCR ~60%, MRD negativity ≥60%, PFS not reached (superior to VRd)More infusion time initially; improved depth of response
KRdCarfilzomib (K) + lenalidomide (R) + dexamethasone (d)Transplant-eligible and fit patients2nd-gen PI + IMiD + steroidFORTE, ENDURANCEORR >90%, CR ~40%, MRD negativity ~30–50%Less neuropathy; more cardiovascular risk than bortezomib
Dara-KRdDaratumumab + KRdTransplant-eligible (under investigation)Anti-CD38 + PI + IMiD + steroidMASTER, MANHATTANMRD negativity >80% in MASTER trialDeepest responses to date; still investigational in frontline
VTDBortezomib + thalidomide + dexamethasoneTransplant-eligible (Europe)PI + IMiD + steroidIFM studiesORR ~85%, CR ~25%, median PFS ~30–35 monthsThalidomide: more neurotoxicity; replaced by lenalidomide in many settings
Dara-VTDDaratumumab + VTDTransplant-eligible (Europe)Anti-CD38 + PI + IMiD + steroidCASSIOPEIAsCR 39% vs. 26%; MRD negativity 64% with Dara-VTDApproved in Europe as standard frontline for transplant-eligible patients
RdLenalidomide + dexamethasoneTransplant-ineligible, frail, or elderly patientsIMiD + steroidFIRST (MM-020)PFS ~26 months (continuous Rd)Better tolerated than triplets; standard for frail or elderly
Dara-RdDaratumumab + RdTransplant-ineligible (standard of care)Anti-CD38 + IMiD + steroidMAIAORR ~93%, MRD negativity >30%, PFS not reached (5+ years)Superior to Rd; preferred triplet for non-transplant-eligible patients
VMPBortezomib + melphalan + prednisoneTransplant-ineligible (Europe, elderly)PI + alkylator + steroidVISTAPFS ~24 monthsUsed mostly in countries without lenalidomide access
Dara-VMPDaratumumab + VMPTransplant-ineligibleAnti-CD38 + PI + alkylator + steroidALCYONEsCR ~43%, MRD negativity ~27%, median PFS ~36 monthsApproved for elderly/NTI; limited in U.S. use
Ixazomib-RdIxazomib (oral PI) + RdTransplant-ineligible, elderly patientsOral PI + IMiD + steroidTOURMALINE-MM2 (did not meet PFS endpoint)Slightly better PFS than Rd; not statistically significantAll-oral regimen; more convenient but less effective than Dara-Rd
PI = proteasome inhibitor; IMiD = immunomodulatory drug; mAb = monoclonal antibody; NTI = non-transplant-eligible; sCR = stringent complete response; MRD = minimal residual disease; PFS = progression-free survival; ORR = overall response rate.
Table 11. Maintenance and consolidation therapies in multiple myeloma.
Table 11. Maintenance and consolidation therapies in multiple myeloma.
Therapy TypeRegimen/DrugPatient PopulationMechanism of ActionKey TrialsOutcomes (PFS/OS)Special Considerations
MaintenanceLenalidomidePost-ASCT; also in non-transplant patientsIMiD; enhances T-cell/NK function, anti-angiogenicCALGB 100104, IFM 2005-02, Myeloma XIMedian PFS benefit: +18–24 months; OS benefit confirmedMost widely used; risk of secondary malignancies (SPMs ~7–9%)
MaintenanceBortezomib (every 2 weeks SC)High-risk cytogenetics (e.g., del(17p), t(4;14))Proteasome inhibition → apoptosisHOVON-65/GMMG-HD4Improved PFS and OS in high-risk MMBetter tolerability SC vs. IV; neuropathy risk
MaintenanceIxazomib (oral)Post-ASCT, especially for convenienceOral proteasome inhibitorTOURMALINE-MM3/MM4PFS improved (~26.5 vs. 21.3 months); no OS benefit yetAll-oral; fewer logistics; mild side effect profile
MaintenanceDaratumumab (monotherapy or with IMiD)Emerging; studied in high-risk and standard-riskAnti-CD38 monoclonal antibodyCASSIOPEIA part 2, AURIGA (ongoing)Early data promising; deepens MRD negativitySC formulation preferred; infusion reactions mainly in induction
MaintenanceCombination (lenalidomide + bortezomib)High-risk patients post-ASCTDual targeting: IMiD + PIFORTE, EMN02/HO95Better MRD negativity; improved PFS in high-risk patientsConsidered for double-hit or high-risk cytogenetics
ConsolidationVRd (2–4 cycles post-ASCT)Transplant-eligible, especially with suboptimal responsePI + IMiD + steroidIFM/DFCI 2009, EMN02/HO95Improved depth of response; modest PFS benefitUseful if no CR/sCR post-ASCT
ConsolidationKRd (carfilzomib-based)Post-ASCT, high-risk or standard-risk2nd-gen PI + IMiD + steroidFORTE (KRd vs. no consolidation)Higher MRD negativity, PFS advantageBetter tolerated in younger/fit patients
ConsolidationDara-VTd/Dara-KRdUnder investigation post-ASCTAnti-CD38-based combinationsGMMG-CONCEPT, MASTER, MANHATTANDeep MRD negativity (>80% in MASTER)Still investigational; tailored to MRD-guided therapy
Maintenance (SMM)Lenalidomide ± dexamethasoneHigh-risk smoldering MMDelays progression to active MMECOG E3A06, ASCENT-SMM3-year PFS ~90% vs. ~60% (control)For selected high-risk SMM patients under clinical trial protocols
Maintenance therapy: low-dose, long-term treatment post-induction/ASCT to prolong remission; consolidation therapy: short-term, intensive treatment post-ASCT to deepen response; ASCT = autologous stem cell transplant; PFS = progression-free survival; OS = overall survival; MRD = minimal residual disease; PI = proteasome inhibitor; IMiD = immunomodulatory drug.
Table 12. Comparison of management strategies in standard-risk vs. high-risk multiple myeloma.
Table 12. Comparison of management strategies in standard-risk vs. high-risk multiple myeloma.
AspectStandard-Risk MMHigh-Risk MM
DefinitionAbsence of high-risk cytogenetic features or clinical markersPresence of high-risk cytogenetics (e.g., del(17p), t(4;14), t(14;16), gain 1q), or double-hit MM (e.g., biallelic TP53)
Common cytogenetic featuresTrisomies (hyperdiploid), t(11;14), no 1q gaindel(17p), t(4;14), t(14;16), gain(1q21), del(1p), complex karyotype, TP53 mutation
Initial risk assessment toolsFISH, R-ISSR2-ISS, NGS, GEP, FISH
Frontline induction therapyTriplet regimens (e.g., VRd, Dara-VRd)Intensified regimens (e.g., KRd, Dara-KRd, quadruplets)
Stem cell transplantSingle ASCTEarly ASCT recommended; tandem ASCT considered in some high-risk cases
Consolidation therapyOptional; may use short course of VRd post-ASCTRecommended (e.g., KRd, Dara-KRd) to deepen response
Maintenance therapyLenalidomide monotherapyBortezomib-based (e.g., bortezomib ± lenalidomide); consider dual-agent maintenance
MRD monitoringMRD monitoring used, but not always guiding therapyEssential; MRD status guides continuation or intensification
Treatment goalsDurable remission; manageable toxicityAchieve and sustain MRD negativity; delay clonal evolution
Role of clinical trialsOptional; used for access to novel agentsStrongly encouraged to access cutting-edge immunotherapy/targeted regimens
Preferred immunotherapy optionsAnti-CD38 mAbs (e.g., daratumumab) in standard regimensEarly use of immunotherapy (e.g., Dara, BiTEs, CAR-T in trials)
Relapse managementSequential therapy (PI/IMiD/monoclonal)Early use of novel agents (e.g., selinexor, CAR-T, bispecifics)
PrognosisMedian PFS > 5 years; OS > 8–10 yearsMedian PFS ~2–3 years; OS ~4–6 years (may vary by subtype)
Monitoring frequencyRoutine labs every 1–3 monthsMore frequent labs, imaging, and MRD assessments
Emerging approachesMRD-guided de-escalation (investigational)MRD-guided intensification, genomics-based precision therapy
Table 13. Comparison of anti-CD38 and anti-SLAMF7 monoclonal antibodies in multiple myeloma management.
Table 13. Comparison of anti-CD38 and anti-SLAMF7 monoclonal antibodies in multiple myeloma management.
FeatureDaratumumabIsatuximabElotuzumab
Target AntigenCD38CD38SLAMF7 (CS1)
Type of AntibodyFully human IgG1κChimeric IgG1Humanized IgG1
Mechanism of Action- Direct apoptosis (crosslinking)
- CDC, ADCC, ADCP
- Immune modulation (Treg depletion)
- Direct apoptosis (even without crosslinking)
- ADCC, CDC, ADCP
- Inhibits CD38 ectoenzyme
- Enhances NK cell activation via SLAMF7
- ADCC (no direct cytotoxicity)
FDA Approval (MM)Yes (monotherapy and in combinations, frontline and relapsed)Yes (in combination regimens for relapsed MM)Yes (only in combination with IMiDs for relapsed MM)
Common Combinations- Dara-Rd (MAIA)
- Dara-VRd (GRIFFIN)
- Dara-KRd
- Dara-Pd (APOLLO)
- Isa-Rd (IKEMA)
- Isa-Kd (IKEMA)
- Elo-Rd (ELOQUENT-2)
- Elo-Pd (ELOQUENT-3)
Line of Use- Frontline (NTI and transplant-eligible)
- Relapsed/refractory (RRMM)
- RRMM (second line and beyond)- RRMM (second line and beyond, not approved for frontline)
Route of AdministrationIV and SC (subcutaneous)IV onlyIV only
Infusion TimeIV: ~3–6 h; SC: ~5 minIV: ~3 hIV: ~2–3 h
Infusion Reactions~40% (mostly first dose); lower with SC~35% first-dose reactions~10–15% (generally mild)
Effect on MRDHigh MRD negativity when combined with IMiDs/PIComparable MRD negativity in IKEMA studyNot MRD-focused; less depth of response
Efficacy in High-Risk MMActive; used in high-risk combinationsActive; included in some high-risk trialsLess data in high-risk subsets
Unique Features- SC formulation reduces infusion burden
- Depletes Tregs and Bregs
- Inhibits CD38 enzymatic function
- More potent direct killing
- Synergizes with NK cells; no single-agent activity
Adverse EffectsCytopenias, infections, IRRs, fatigueSimilar to daratumumabFatigue, diarrhea, cough, IRRs
Use with IMiDs or PIsWorks well with both IMiDs and PIsPrimarily used with IMiDsOnly approved with IMiDs (lenalidomide, pomalidomide)
Trial Highlights- MAIA (Dara-Rd)
- GRIFFIN (Dara-VRd)
- CASSIOPEIA (Dara-VTd)
- IKEMA (Isa-Kd)
- ICARIA-MM (Isa-Pd)
- ELOQUENT-2 (Elo-Rd)
- ELOQUENT-3 (Elo-Pd)
Table 14. Comparison of idecabtagene vicleucel (ide-cel) vs. ciltacabtagene autoleucel (cilta-cel) in multiple myeloma.
Table 14. Comparison of idecabtagene vicleucel (ide-cel) vs. ciltacabtagene autoleucel (cilta-cel) in multiple myeloma.
FeatureIdecabtagene Vicleucel (ide-cel; Abecma®)Ciltacabtagene Autoleucel (cilta-cel; Carvykti®)
Target antigenB-cell maturation antigen (BCMA)BCMA
CAR structureSingle-chain variable fragment (scFv)Two BCMA-binding domains (dual-epitope targeting scFvs)
CAR-T-cell generationLentiviral vector, autologous T cellsLentiviral vector, autologous T cells
FDA approvalMarch 2021February 2022
IndicationTriple-class refractory MM (≥4 prior lines)Triple-class refractory MM (≥4 prior lines)
Key clinical trialKarMMa (Phase 2)CARTITUDE-1 (Phase 1b/2)
Overall response rate (ORR)~73%~98%
Stringent complete response (sCR)~33%~80%
Median progression-free survival~8.8 months (median follow-up 13.3 months)Not reached at 27 months; PFS at 2 years: ~61%
Median overall survival~24.8 monthsNot reached at 2 years; OS ~74%
Time to response~1 month~1 month
Time from apheresis to infusion~4–5 weeks~4–6 weeks
Toxicity—CRS (cytokine release syndrome)~84% (mostly Grade 1–2); Grade ≥3 in ~5%~95%; mostly Grade 1–2; Grade ≥3 in ~4%
Toxicity—ICANS (neurotoxicity)~18% (Grade ≥3: 3%)~21% (Grade ≥3: 10%)
Unique featuresFirst FDA-approved BCMA CAR-T; robust data in heavily pretreated patientsDual-epitope binding = higher avidity; very deep and durable responses
Dosing regimenSingle infusion of 150–450 × 106 CAR+ T cellsSingle infusion of 0.75 × 106 CAR+ T cells/kg
Emerging directionsUsed in KarMMa-3 (earlier lines); potential combinationsCARTITUDE-4: early relapse; being explored for frontline high-risk MM
Table 15. Grading of cytokine release syndrome (CRS).
Table 15. Grading of cytokine release syndrome (CRS).
GradeFever ≥ 38 °CHypotensionHypoxia
Grade 1PresentNoneNone
Grade 2PresentDoes not require vasopressors; responds to fluidsRequires low-flow oxygen (nasal cannula ≤6 L/min)
Grade 3PresentRequires vasopressors or multiple bolusesRequires high-flow oxygen, CPAP, or BiPAP
Grade 4PresentRequires multiple vasopressors ± ventilatory supportRequires mechanical ventilation
Table 16. Grading of immune effector cell-associated neurotoxicity syndrome (ICANS).
Table 16. Grading of immune effector cell-associated neurotoxicity syndrome (ICANS).
GradeICE ScoreLevel of ConsciousnessSeizureMotor FindingsCerebral Edema
Grade 17–9Awaken spontaneouslyNoneNoneNone
Grade 23–6Awaken to voiceNoneNoneNone
Grade 30–2Arousable with tactile stimulationAny seizureNoneFocal/local edema in neuroimaging
Grade 40Unarousable or obtunded; stupor or comaLife-threatening prolonged seizureDeep focal motor weakness, such as hemiparesis or paraparesisDiffuse cerebral edema in neuroimaging; decerebrate or decorticate
ICE score components (maximum 10 points): orientation to year, month, city, hospital = 4; naming 3 objects = 3; following simple command = 1; writing a standard sentence = 1; attention (counting backwards from 100 by 10) = 1.
Table 17. Investigational and available cancer vaccines in multiple myeloma.
Table 17. Investigational and available cancer vaccines in multiple myeloma.
Vaccine Name/PlatformAntigen Target(s)Vaccine TypeClinical Phase/StatusMechanism of ActionKey Studies/Notes
PVX-410XBP1, CD138, CS1 (SLAMF7)Peptide-based multi-epitope vaccinePhase I (completed)Induces CD8+ T-cell responses against MM-associated antigensSafe and immunogenic in SMM; studied alone and with lenalidomide (NCT01718899)
Bcl-2 family peptide vaccineBcl-2, Bcl-XL, Mcl-1Peptide-based vaccinePhase IStimulates cytotoxic T cells targeting anti-apoptotic proteinsCombined with bortezomib; immunogenic with minimal toxicity
MAGE-A3 peptide vaccineMAGE-A3Tumor antigen-specific peptide vaccineEarly phase/exploratoryTargets cancer–testis antigen expressed in MM and other malignanciesUsed in pilot studies and as a possible post-ASCT strategy
Idiotype (Id) protein vaccinePatient-specific paraprotein (Id)Personalized protein vaccinePhase I/II (historical interest)Induces helper and cytotoxic T-cell responses against myeloma idiotypeLimited efficacy; used in post-ASCT settings; declining interest
DC/MM fusion vaccineWhole-tumor antigens from patient myeloma cellsDendritic cell (DC)-tumor fusion vaccinePhase I/IIEnhances antigen presentation by fusing DCs with autologous MM cellsShown to induce tumor-specific immunity post-ASCT (NCT02728102, NCT01067287)
Neoantigen-based vaccinesPatient-specific neoantigensPersonalized peptide/RNA vaccinePreclinical/early phaseStimulates T cells against mutation-specific targets unique to each patientRequires whole-exome sequencing and epitope prediction
Gp96-Ig vaccine (ImPACT®)Broad tumor antigen spectrumHeat shock protein-based chaperone vaccinePreclinical/exploratoryCross-presents multiple antigens via HLA to CD8+ T cellsUsed in MM and other hematologic malignancies
mRNA-based vaccines (exploratory)Personalized neoantigens or MM-associated antigensmRNA encoding tumor antigensPreclinicalInduces cellular immunity via in vivo expression of encoded antigensEmerging platform post-COVID-19 vaccine success
WT1 peptide vaccineWilms’ tumor 1 (WT1)Peptide vaccinePilot/early phaseTargets overexpressed WT1 protein in MMStudied in combination with other immunotherapies
NY-ESO-1 vaccineNY-ESO-1Cancer–testis antigen peptide vaccineEarly phaseInduces T-cell response against highly immunogenic antigenUnder study in hematologic and solid tumors
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sutanto, H.; Romadhon, P.Z.; Fatmawati, V.R.; Waitupu, A.; Ansharullah, B.A.; Rachma, B.; Elisa, E.; Pratiwi, L.; Adytia, G.J. Multiple Myeloma and Precursor Plasma Cell Disorders: From Emerging Driver Mutations to Current and Future Therapeutic Strategies. Hemato 2025, 6, 29. https://doi.org/10.3390/hemato6030029

AMA Style

Sutanto H, Romadhon PZ, Fatmawati VR, Waitupu A, Ansharullah BA, Rachma B, Elisa E, Pratiwi L, Adytia GJ. Multiple Myeloma and Precursor Plasma Cell Disorders: From Emerging Driver Mutations to Current and Future Therapeutic Strategies. Hemato. 2025; 6(3):29. https://doi.org/10.3390/hemato6030029

Chicago/Turabian Style

Sutanto, Henry, Pradana Zaky Romadhon, Vembi Rizky Fatmawati, Alief Waitupu, Bagus Aditya Ansharullah, Betty Rachma, Elisa Elisa, Laras Pratiwi, and Galih Januar Adytia. 2025. "Multiple Myeloma and Precursor Plasma Cell Disorders: From Emerging Driver Mutations to Current and Future Therapeutic Strategies" Hemato 6, no. 3: 29. https://doi.org/10.3390/hemato6030029

APA Style

Sutanto, H., Romadhon, P. Z., Fatmawati, V. R., Waitupu, A., Ansharullah, B. A., Rachma, B., Elisa, E., Pratiwi, L., & Adytia, G. J. (2025). Multiple Myeloma and Precursor Plasma Cell Disorders: From Emerging Driver Mutations to Current and Future Therapeutic Strategies. Hemato, 6(3), 29. https://doi.org/10.3390/hemato6030029

Article Metrics

Back to TopTop