Next Article in Journal
Expression of CK17 and SOX2 in Vulvar Intraepithelial Neoplasia: A Comprehensive Analysis of 150 Vulvar Lesions
Previous Article in Journal
Multiple Myeloma Cells with Increased Proteasomal and ER Stress Are Hypersensitive to ATX-101, an Experimental Peptide Drug Targeting PCNA
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Unraveling the Genetic Heterogeneity of Acute Lymphoblastic Leukemia Based on NGS Applications

by
Valentina Ramírez Maldonado
1,
Josgrey Navas Acosta
1,
Iván Maldonado Marcos
1,
Ángela Villaverde Ramiro
1,
Alberto Hernández-Sánchez
1,2,
Jesús M. Hernández Rivas
1,2 and
Rocío Benito Sánchez
1,*
1
Centro de Investigación del Cáncer, IBMCC, CSIC, Universidad de Salamanca, IBSAL (Instituto de Investigación Biomédica de Salamanca) Campus, Miguel de Unamuno, 37007 Salamanca, Spain
2
Servicio de Hematología, Complejo Asistencial Universitario de Salamanca, 37007 Salamanca, Spain
*
Author to whom correspondence should be addressed.
Cancers 2024, 16(23), 3965; https://doi.org/10.3390/cancers16233965
Submission received: 17 October 2024 / Revised: 17 November 2024 / Accepted: 22 November 2024 / Published: 26 November 2024
(This article belongs to the Special Issue Algorithms and Data Analysis of High Throughput Sequencing in Cancers)

Simple Summary

Acute lymphoblastic leukemia (ALL) remains the most frequent cancer among pediatric patients and accounts for about a quarter of adult leukemias. Understanding the biological processes through which leukemogenesis originates and is maintained remains a challenge. Advances in different omics technologies have broadened our understanding and have enabled better diagnosis and management of patients by enabling the identification of biomarkers with prognostic impact, including genomic, transcriptomic, and epigenomic markers. Therefore, the aim of this review is to compare, summarize, and highlight some of the recent findings in acute lymphoblastic leukemia based on NGS applications.

Abstract

Acute lymphoblastic leukemia (ALL) is a hematological neoplasm characterized by the clonal expansion of abnormal lymphoid precursors in bone marrow, which leads to alterations in the processes of cell differentiation and maturation as a consequence of genetic alterations. The integration of conventional methods, such as cytogenetics and immunophenotyping, and next-generation sequencing (NGS) has led to significant improvements at diagnosis and patient stratification; this has also allowed the discovery of several novel molecular entities with specific genetic variants that may drive the processes of leukemogenesis. Nevertheless, the understanding of the process of leukemogenesis remains a challenge since this disease persists as the most frequent cancer in children; it accounts for approximately one-quarter of adult acute leukemias, and the patient management may take into consideration the high intra- and inter-tumor heterogeneity and the relapse risk due to the various molecular events that can occur during clonal evolution. Some germline variants have been identified as risk factors or have been found to be related to the response to treatment. Therefore, better knowledge of the genetic alterations in B-ALL will have a prognostic impact from the perspective of personalized medicine. This review aims to compare, synthesize, and highlight recent findings concerning ALL obtained through NGS that have led to a better understanding of new molecular subtypes based on immunophenotypic characteristics, mutational profiles, and expression profiles.

1. Introduction

Acute lymphoblastic leukemia (ALL) is a hematological malignancy characterized by the clonal expansion of abnormal lymphoid precursors in the bone marrow, as a consequence of alterations in cell differentiation and maturation processes through the acquisition of aneuploidies, chromosomal rearrangements, and driver mutations in genes related to lymphoid cells development [1,2]. It can affect the lymph nodes, thymus, spleen, testicles, and central nervous system, and it may be derived from precursors of both T (T-ALL) or B (B-ALL) lineages [3,4]. B-ALL is the most common phenotype in both pediatric and adult patients, representing 75% and 85% of cases, respectively [5,6].
Although ALL is considered an orphan disease given that it has an age-standardized incidence rate (of 0.4 to 2.12 per 100,000 people per year) [7,8], this neoplasm still represents public health since around 6000 new diagnoses are reported each year [9]. In addition, it remains the most common childhood neoplasm, representing between 14% and 19% of all childhood cancers. It is also the second most common acute leukemia in adults [10]. Despite advances in diagnosis and the use of targeted therapies that have made it possible to achieve 5-year event-free survival rates of around 80% and 5-year survival rates approaching 90% in pediatric patients, as well as rates between 75 and 80% in adolescents and young adults (AYAs) [11], about 15% of these patients may present relapsed/refractory (R/R) disease [12,13]. Regarding adults, long-term survival decreases with age, and 5-year overall survival is estimated to be around 40% in those older than 40 years [8,14]. Therefore, a better understanding of the biological mechanisms and identifying biomarkers that can be associated with leukemogenesis remains a challenge.
ALL diagnosis requires the presence of 20% or more lymphoblasts in the bone marrow or peripheral blood, based on morphological assessment. Classification must be performed using cell morphology, immunohistochemistry, and immunophenotyping through flow cytometry. Karyotyping is still considered the “gold standard”, given that the majority of chromosomal rearrangements or aneuploidies are well characterized [6,15], may represent a strong independent prognostic factor [16], and can be used for risk assignment for treatment and therapy selection [3,17]. Nevertheless, approximately 25% of ALL cases may have a normal karyotype that could harbor high rates of submicroscopic aberrations, or cytogenetic analysis could fail [18,19,20]. In addition, around 25% to 40% of ALL patients display defined genetic abnormalities that cannot be identified by these methods [21]. In this sense, other approaches of molecular biology with higher sensitivity and specificity such as multiplex ligation-dependent probe amplification (MLPA), chromosomal microarrays (CMA), and comparative genomic hybridization (aCGH), may be also used in clinical practice [22]. Moreover, the development and implementation of next-generation sequencing (NGS) technologies for whole genome sequencing (WGS), whole exome sequencing (WES), whole transcriptome analysis (WTS), target sequencing (TGS), and epigenetic analysis have resulted in the identification of new genetic variants including point mutations, intrachromosomal amplifications, or copy number variants (CNVs) in genes such PAX5, TP53, IKZF1, CDKN2A/2B, and RB1 in B-ALL, as well as NOTCH1, FBXW7, and WT1 in T-ALL [23], to assess specific altered signaling pathways such as RAS or PI3K/AKT/mTOR [24], the association the presence of polymorphisms with toxicity or response to treatment in genes such as MTHFR (rs1801133) or TPMT [25,26], and also different epigenetic signatures according to phenotype and genetic subtype [27,28]. These results have allowed for a better understanding of leukemogenesis, thus providing different classifications to establish genetic subtypes associated with prognoses such as those proposed by the World Health Organization (WHO-HAEM5) or the International Consensus Classification of Myeloid Neoplasm and Acute Leukemias (ICC). In addition, the following guidelines for the management of ALL have been provided: 2024 ELN recommendations from a European expert panel for Philadelphia-negative, Philadelphia-positive, or Philadelphia-like subgroups [29] and NCCN (2024). This translates into better ALL patient management when the molecular characteristics and clinical features are integrated. In this context, this review aims to compare, synthesize, and highlight the recent insights into ALL based on NGS analysis using a narrative framework (Panel A), an opportunity for personalized medicine.
Panel A. Search strategy and selection criteria
Keywords such as “acute lymphoblastic Leukemia”, “high-throughput sequencing”, “massive sequencing”, “NGS”, “ALL risk factor”, “molecular classification”, “polymorphism”, “omics data”, and their combinations were used. Tags such as [tiab][res] or [intext: intittle] and quotation marks as search operators and (*) were used as wildcard characters or to narrow the search.

The inclusion criteria were established using the following: 1. subject area and sub-areas (medicine, hematology, oncohematology, genetics), 2. type of study and publication (clinical and in vitro studies, clinical trials, observational studies, clinical practice guidelines), 3. period of analysis and language (the initial search was conducted between 2017 and 2023, and older references were also used where appropriate; only English or Spanish articles were included), and 4. availability: only open access documents were included. References include original articles, systematic reviews, and meta-analyses.

The information retrieval was performed using Health in Science Descriptors (Medical Subject Headings of the National Library of Medicine of the USA-MeSH) and Hierarchical Code/thesauri, as well as databases such as Elsevier-Science Direct, Web of Science (Web of Knowledge-WoK), Scopus, and PubMed.

2. Next-Generation Sequencing (NGS) Applications in ALL and New Insights

In recent years, the genomic analysis of ALL in clinical practice has been performed using standard techniques that allow for the detection of the most frequent alterations defining molecular subtypes, such as ploidy alterations (hyperdiploidy or hypodiploidy), intrachromosomal amplifications, or gene fusions (BCR::ABL1, ETV6::RUNX1, KMT2Ar) (Figure 1).
However, these methods are limited in the assessment of other cryptic rearrangements that were recently proposed for B-ALL, such as DUX4, EPOR, ZNF384 and MYC rearrangements, and single-point mutations (PAX P80R or IKZF1 N159Y) [30], as well as T-ALL subtypes defined by TAL-LMO or LYL1 in association with gene expression signatures [31,32]. In addition, copy number variants (CNVs) and phenocopies cannot be detected with these methods. For this reason, NGS methods have been proposed to improve the characterization and detection of biomarkers that allow for better certification and management of ALL patients in the context of personalized medicine
The advances in NGS have revolutionized our understanding of leukemogenesis through the characterization of comprehensive genomic, transcriptomic, and epigenomic profiles [33,34], which provide unprecedented insight into the genetic and molecular heterogeneity of ALL. In addition, these technologies offer a powerful method for detecting a wide range of genetic variants, SNPs, insertions, deletions, and structural rearrangements and enable the identification of numerous genetic subtypes of ALL with prognostic value (Table 1). On the other hand, although the implementation of these technologies in clinical practice is promising because they allows diagnoses to be refined and could contribute to the personalization of treatment strategies in ALL patients, some limitations must be taken into consideration (Table 2).
NGS-based methods currently display variable sensitivity in detecting copy number variations (CNVs) and chromosomal rearrangements, limiting their ability to offer a fully comprehensive view of genomic alterations. As bioinformatics tools for analyzing CNV data continue to evolve, it remains essential to use NGS-based CNV analysis in combination with well-established cytogenetic methods, such as karyotyping, fluorescence in situ hybridization (FISH), SNP arrays, multiplex ligation-dependent probe amplification (MLPA), or optical genome mapping (OGM). These additional techniques provide complementary insights that address the limitations of NGS and help ensure accurate structural variant detection.
OGM is showing promise for future clinical applications, with emerging evidence supporting its utility in practice. The International Consortium for Optical Genome Mapping in Hematologic Malignancies has recently demonstrated a framework for clinical implementation, highlighting the technology’s potential to enhance diagnostic precision in hematologic malignancies [51,52]
Due to the complexity of CNV analysis, where detection sensitivity can vary based on sequencing strategies and variant sizes, NGS-based CNV detection should employ a multi-algorithmic approach. Comparing results across different bioinformatics tools, particularly when using targeted sequencing panels, can greatly enhance the accuracy of CNV identification [53,54]

3. New Insights into ALL Genomic Profiling and Recent Classification Based on Molecular Features

From its early stages, the classification of acute lymphoblastic leukemia (ALL) has undergone significant evolution. The first ALL classification was proposed by the French–American–British group in 1976 based on morphological and cytochemical criteria. This foundational system was later expanded by the WHO to include cytogenetic features [55]. With the advent of flow cytometry and high-throughput sequencing technologies, newer and more precise classifications have emerged. In 2022, the International Agency for Research on Cancer (IARC) released the latest version of the WHO classification of hematolymphoid tumors (WHO-HAEM5). This classification system establishes hierarchical categories that include family/class and entity/type and proposes thirteen categories for B-lymphocytic leukemia/lymphoma and two for T-lymphocytic leukemia/lymphoma [55,56]. Additionally, other groups, such as the German Multicenter Study group for Adult ALL (GMALL), the European Group for Immunological Classification of Leukemias (EGIL), the Munich Leukemia Laboratory (MLL), and the International Consensus Classification (ICC) of Myeloid Neoplasm and Acute Leukemias, have suggested incorporating additional molecular subtypes. These recommendations are based on advanced NGS applications such as WES or WTS with gene expression profile (GEP) analysis to identify phenocopies and novel fusion genes (Table 3 and Figure 2). These evolving classifications reflect the continuous advancements in genomic technologies and their profound impact on understanding and treating ALL.

4. Molecular Landscape of New and Emerging Genetic Subtypes of B-ALL

4.1. New Insights in Philadelphia ALL Patients (BCR::ABL1-Positive)

Although Philadelphia-positive ALL (Ph+) patients are typically characterized by BCR::ABL1 fusion, the use of NGS has demonstrated that the genomic and transcriptomic profiles can be heterogeneous. Kim et al. (2020) [59] proposed two subtypes according to the BCR::ABL1 isoform, and the cell type from which it came: the long isoform (p210) originating from a hematopoietic stem cell (HSC) that could generate a multi-lineage (multi-Ph) phenotype; and patients with short isoform (p190) and uni-lineage phenotype, from a B-cell progenitor [59]. However, these results were controversial, as recent follow-up work on residual disease shows that the p190 isoform can also be maintained in HSC with a typical B-cell phenotype [60].
Recently, Jaeseung C. Kim et al. (2023) [61] employed WES, WGS, scRNA-seq, and ATAC-seq and proposed new subtypes with prognostic value in Ph+ ALL patients. Each subtype was defined by cooperating events in genes that regulate B-cell development and not by the cell of origin where the event occurs. C1 (from an early B-cell progenitor, “Early-Pro”), C2 (from an intermediate B-cell progenitor, “Inter-Pro”), C3 (late B-cell progenitor, “Late-Pro”). C1 presented higher leukocyte counts at diagnosis and was characterized by aberrant expression of stem cell and myeloid genes, such as KIT, CSF3R, and MECOM, as well as an increased expression of genes related to innate immunity, such as TNFα. Group C2 showed differential expression of both myeloid (CSF2RA and CSF1R) and lymphoid (MS4A1/CD20 and IL7R) genes, as well as genes related to the ERK signaling and unfolded protein response (UPR) pathways. Group C3 showed the highest expression of genes associated with B-cell differentiation, such as IL7R, MS4A1, BACH2, and TCL1A, and cell cycle regulatory genes. Analysis of IGH heavy-chain expression was similar in all groups, while the expression of rearranged light chains (IGK or IGL) was more frequent in the C3 group, which also showed higher expression of B-cell factors such as EBF1 and PAX5, suggesting a later differentiation block compared with the other groups [61].
Another relevant fact among Ph+ leukemias is the IKZF1 deletions and other cooperative alterations in genes related to B-cell differentiation such as PAX5, CDKN2A/B and EBF1, in addition to other genes such as BTG1, RB1, EBF1, MEF2C, RUNX1, SLX4IP, HBS1L, TSC22D1, FOXP1, LEF1, ETV6, TCF3 and TCF4 [62]. Jaeseung C. Kim et al. (2023) conducted a WGS analysis and described new recurrent alterations in CBWD2, GPN3/FAM216A, PRKAR2B (without monosomy 7), MAP3K2, MIR181A1HG/MIR181B1, and FOXP1. C1 leukemias present a higher proportion of alterations, such as monosomy 7 or 1q gain, and HBS1L deletions; EBF1 and RUNX1 variants are restricted to this group. However, C2 and C3 present a higher number of alterations in genes such as PAX5, CDKN2A, and SLX4IP, as well as hyperdiploidy. In addition, although IKZF1 alterations were present in all subtypes, biallelic losses and the dominant negative Ik6 deletion of IKZF1 were more frequent in C2, suggesting that complete inactivation IKZF1 drives the Inter-Pro subtype. The C3 group had higher frequencies of PAX5 deletions and homozygous deletions of CDKN2A/B and RB1, which was consistent with a posterior block in B-cell differentiation.
Regarding the immunophenotype and molecular response, differences were also found between the groups. C1 presented myeloid antigens, CD13, CD33, and cytoplasmic myeloperoxidase (cyMPO), CD7, and the highest expression of CD34, while the C3 group displayed the highest levels of CD10, CD19, and CD20. Fewer C1 and C2 patients achieved a major molecular response (MMR, log reduction ≥3, transcript ≤0.1%) compared with C3 patients, who had a deep molecular response (log reduction ≥4); consequently, they had higher overall survival (OS) and event-free survival (EFS) rates. On the other hand, the authors also propose that prognostic staging should take the occurrence of mutations inside or outside the BCR::ABL kinase domain into account. C1 patients exhibit an enrichment of BCR-ABL1-signaling pathways and its target STAT5, as well as increased phosphorylation of STAT5. This has been described as a mechanism of resistance to TKI in the absence of kinase domain mutations. In contrast, Lorenz Bastian et al. (2024) [63] proposed two main clusters within ALL Ph+ based on transcriptomic and genomic profiling, as well as ATAC-Seq analysis. The authors found 331 differentially expressed genes that allowed for the establishment of complex transcriptomic signatures for the two main and four subclusters: multilineage (delHBS1L or del7) and lymphoid (homozygous IKZF1 deletion, CDKN2A/PAX5, and/or hyperdiploidy with strong enrichment for the Hippo and Hedgehog pathways, respectively). HBS1L deletion harbored the same breakpoints in all samples (chr6:135,044,863-135,116,862; GRCh38hg38), including the promoter and exons 1 to 2. Patients showed an increased HBS1L gene expression but reduced usage of exon 1 to 3. Long-read RNA-Seq confirmed a previously undescribed HBS1L transcript (HBS1Lalt) in delHBS1L and del7 cases initiated from a putative transcription start side in HBS1L intron 3 [63].
Taken together, the results suggest DNA- or RNA-based risk stratification may have clinical utility. In addition, studies should be expanded to conclude on the association of the phenotype with other clinical variables. Jaeseung C. Kim et al. (2023) indicate that subtypes are not associated with age, sex, or the BCR-ABL1 isoform (p190 or p210). In comparison, Lorenz Bastian et al. (2024) found a predominance of the CDKN2A/PAX5 subgroup in pediatric patients despite the equal distribution of multilineage and lymphoid cases in children and AYAs. Adult patients were more frequently classified as multilineage with an increase in del7 cases. In addition, patients in the multilineage group more frequently presented major breaks in BCR::ABL1, with a significant increase in BCR exon14 involvement in delHBS1L. Moreover, white blood cell counts at diagnosis also differed between the new BCR::ABL1 groups (highest in delHBS1L and CDKN2A/PAX5), indicating that the new group definitions also represent distinct clinical profiles.

4.2. Phenocopies in B-ALL

4.2.1. BCR::ABL1-like (Ph-like)

This subgroup presents a gene expression profile similar to that of Philadelphia-positive patients; it occurs at all ages, although it is less common in children (<10%), and its prevalence increases to 30% in adolescents. Comprehensive NGS (WTS and WGS simultaneously) is the best approach to identifying this subtype. It is commonly characterized by CRLF2 rearrangements (around 50% of cases) and IKZF1 deletions. It is associated with adverse clinical features, such as a high white blood cell count, a high minimal residual disease level after induction therapy, and a high relapse rate [64]. Thus, Arber et al. (2022) highlighted the importance of distinguishing molecular subtypes among BCR::ABL1-like (Philadelphia-like, Ph-like) patients, considering that some patients could benefit from specific therapies [65]. These patients can be grouped as follows: (i) The ABL class, which includes 5′ ABL1 fusions (with CENPC, ETV6, FOXP1, LSM14A, NUP153, NUP214, RANBP2, RCSD1, SFPQ, SNX1, SNX2, SPTNA1, or ZMIZ1), ABL2 fusion (with PAG1, RCSD1, or ZC3HAV1); CSF1R fusions (with MEF2D, SSBP2 or TBL1XR1), PDGFRB fusions (with ATF7IP, EBF1, ETV6, SNX29, SSBP2, TNIP1, ZEB2, or ZMYND8); and PDGFRA::IP1L1 [66,67]. (ii) Altered JAK-STAT signaling, with JAK1, JAK2, IL7R, and SH2B3 mutations or EPOR rearrangements [68,69]. JAK2 gain-of-function mutations may occur in association with CRLF2 overexpression in around 50% of cases. The latter occurs as a consequence of translocation, leading to IGH-CRLF2 juxtaposition or a PAR deletion (PAR1, Xp22.33/Yp11.2) and generating P2RY8-CRLF2 fusion [70,71]. The third grouping is (iii) BCR::ABL1-like not otherwise specified (NOS), with genetic fusions or mutations involving genes such as ETV6, BLNK, DGKH, PTK2B, FLT3, FGFR1, and TYK2. In total, 1% of Ph-like cases harbor an ETV6::NTRK3 fusion that is associated with aggressive ALL [72,73]. On the other hand, CNV analysis of IKZF, CDKN2A/B, PAX5, PAR1, and ERG in BCR::ABL1-like patients has been proposed as it could allow for the identification of IKZF1plus or IKZF1-deleted profiles that may have prognostic significance, and this could guide more personalized therapeutic strategies to improve outcomes in this group [54].

4.2.2. ETV6::RUNX1-like

This genotype occurs in approximately 2–3% of cases in children and in less than 1% of cases in AYAs and adults [74,75]. This subgroup harbors ETV6 aberrations and is characterized by a CD27-positive/CD44-low to -negative immunophenotype and a similar gene expression profile to that of the ETV6::RUNX1 subtype. This genotype may harbor IKZF1, TCF3, ERG, or ARPP21 CNVs [76]; gene fusions such as IKZF1::ETV6, ETV6::ELMO1, TCF3::FLI1, or FUS::ERG [77,78]; and biallelic inactivation of ETV6 by de novo deletion or in pediatric patients that harbor a germline ETV6 mutations with subsequent somatic alterations of the second ETV6 allele. This biallelic ETV6 alteration is central in leukemogenesis [76,79]. On the other hand, Dongfeng Chen and colleagues (2020) proposed a RAG1 signature belonging to the ETV6::RUNX1-like subtype, which shows high RAG1 mRNA levels and over-expression of ETV6 target genes as well as BIRC7, WBP1L, CLIC5, and ANGPTL2 [75]. The prognostic value is unclear. Some studies have shown an average 5-year EFS of 66.7%, 5-year OS > 87%, and an MRD negative at the end of consolidation (EOC) of >86%, suggesting that this subtype may benefit from higher intensity therapy [2,80]. However, Jeha S et al. (2021) suggest that this genotype confers an unfavorable prognosis in children due to the high levels of MRD and worse event-free survival rates [81].

4.2.3. ZNF384r-like (Provisional Entity)

This subtype was characterized using RNA-seq, gene expression profiles (GEPs), and a machine learning classifier (ALLSorts) [82]. This cluster presents a similar ZNF384r expression profile without 5′ canonical changes in ZNF384, including indels such as p.393_397KHNPP>RAG, p.483_484CS>YFEWDPND, and p.481_516>SFQDYCACQ that lead to a disruption in the C-terminal [83]. Likewise, a ZNF362(ZNF384 paralog) rearrangement has been identified in this subgroup. Similarly to ZNF384r leukemias, this entity has a predominant BCP immunophenotype. Myeloid markers (CD13/CD33) are often expressed, CD10 is weakly expressed, and some patients may also have aberrant expression of CD66c [84,85]. The prognostic value is yet unknown.

4.2.4. KMT2Ar-like (Provisional Entity)

This subgroup was proposed by Amy S. Duffield et al. (2023) based on WTS and gene-expression-profiling data from 2041 leukemia samples collected at diagnosis. They found some cases (<1%) with a differential profile between KMT2Ar cases. They proposed that these patients may harbor HOXA fusion. As with ZNF384r-like, the prognostic value is currently unknown [58].

4.3. Novel Gene Fusions in B-ALL

4.3.1. MYC-Rearranged

This rearrangement may occur around 2% to 5% cases B-ALL, and it is rare in T-ALL (1% in children, and it is under-recognized in adults) [86,87]. It is associated with an extremely aggressive syndrome characterized by hyperleukocytosis, rapid neurological progression, and poor response to chemotherapy [88,89]. However, recent studies show that the combination of intensive chemotherapy and rituximab can improve the prognosis, allowing for the cure of about 50% of this subtype [90,91]. Usually, these patients present MYC overexpression due to t(8;14)(q24;q11), which places the MYC gene under the control of strong T-cell-responsive enhancer elements (TCRA or TCRD) [92] or other partner genes, including IGL, IGK, BCL2 (<3%) [16,93,94], and BCL6 (<1%) [56,58]. The prognostic significance varies according to the partner gene. Recent studies suggest that MYC::IGH fusion could confer good prognosis with high-dose methotrexate and cyclophosphamide [16]. In contrast, MYC::BCL2 can be associated with a worse outcome [16].

4.3.2. IGH Fusions, t(v;14q32)

These fusions represent around 5% of all B-ALL patients, and the rate is higher in adults. IGH fusions are usually associated with poor prognosis independent of the partners genes, including BCL2, BCL6, ID4, MYC, CEBPA, CEBPB, IGF2BP1, IGK, EPOR, LHX4, IL-3, and CRLF2 [16,86,95]. Nevertheless, specific cases, such as IGH::BCL2, should be rigorously analyzed since low levels of circulating t(14;18)-positive cells have been found in approximately 50–70% of healthy individuals who never develop ALL [80,81]. IGH::CRLF2 fusion is found almost exclusively in the subset of Ph-like ALL (~50%) [94,96].

4.3.3. TCF3-Rearranged (19p13.3)

This accounts for about 4–6% of all pediatric B-ALL cases and is less frequent in adults (<3%) [2,16,97]. TCF3r is overrepresented in hyperdiploidy and ETV6::RUNX1 ALL cases (~8%) [94], and it may be accompanied by IKZF1 deletions [2]. The phenotype is characterized by moderate CD10 and CD19, strong expression of CD9, low to negative CD34, and at least a partial absence of CD20. This may be associated with PAX5 or VPREB1 deletions or mutations in Ras pathway genes (NRAS, KRAS or PTPN11) [98,99,100]. Although der(19)t(1;19) chromosomes can be detected using conventional cytogenetics or FISH, using other NGS strategies has allowed for the characterization of additional TCF3 partner genes. The most common rearrangement is TCF3::PBX1 fusion involving TCF3 exons 1–16 and exons 3–9 of PBX1. However, in up to 10% of cases, there are alternative break-points [101]. Other partners genes include ZNF384 (12p13) [102], FLI1 (11q24.3), TEF (22q13.2), and HLF (17q22). HLF is not normally transcribed in lymphoid cells, but fusion leads to the leukemic cells becoming similar to stem cells [100,101,103]. In addition, Fischer et al. found an association of TCF3/HLF with PAX5 haploinsufficiency and enrichment in stem cell and myeloid expression signatures [104,105]. Even though it is present in <1% of ALL cases, this genotype was recently included in WHO-HAEM5. In addition, it is associated with poor prognosis, early relapse, and death within two years of diagnosis [106,107]. Another relevant aspect is the consideration that these patients exhibit a high CD19 expression, and immunotherapy, stem cell transplantation, and CAR T-cell therapy may hold promise [108].

4.3.4. DUX4-Rearranged

DUX4r may be present in 3% to 10% of the cases of B-ALL, with a higher proportion of cases in AYAs; it is associated with a favorable prognosis, despite the presence of high MRD levels [109,110]. These cases have a distinct CD2-positive and CD13-, CD34-, and CD38-overexpression immunophenotype [111], as well as a GEP characterized by DUX4 and ERG deregulation. The main rearrangement is a linkage between IGH and DUX4 (4q and 10q D4Z4 repeat domains), but t(4;21)(q35;q22) DUX4/ERG also may occur [111,112]. DUX4 over-expression can be detected using IHC, but according to the diagnosis criteria in WHO-HAEM5, RNA or DNA sequencing must be used to identify other cryptic rearrangements and simultaneous co-lesions, such as NRAS- or KRAS-activating mutations, IKZF1, CDKN2A/B, or ERG deletions or ERG loss of function (ERGalt). ERGalt is generated by aberrant DUX4 binding to intragenic region ERG, which leads to an ERG isoform that retains the DNA-binding and transactivation domains inhibiting the ERG wild-type allele with a dominant negative effect [76,113]. Moreover, genomic studies have revealed that around 50% to 70% of cases display intragenic ERG deletions, a genomic aberration that is practically nonexistent in other B-ALL cases [114].

4.3.5. ZNF384-Rearranged (12p13.31)

This represents 3% of all pediatric B-ALL cases and around 7–10% of cases in AYAs. It has been proposed that its prognostic value is controversial [115]. Ya-Zhen Qin et al. (2021) found significantly higher 3-year relapse-free survival (RFS) and less overall survival (OS) compared with those without [116]. In contrast, Zhu et al. (2023) proposed that patients commonly have a good prognosis, but this might depend on the fusion partner gene [117]. This subtype exhibits abnormal expression myeloid markers CD13 and/or CD33. However, the typical marker of an immature B lineage, CD10, is only weakly expressed in some of them. The most common fusion partners are EP300/ZNF384, CREBBP/ZNF384, and TCF3/ZNF384, but TAF15, EWSR1, ARIDIB, BMP2K, CLTC, NIPBL, SMARCA2, and SYNRG have also been reported [85] [115,117,118]. ZNF384r carries other co-lesions, such as Ras pathway gene mutations; FLT3 overexpression; CLCF1 and BTLA upregulation; JAK-STAT pathway activation due to GATA3, CEBPA, and CEBPB over-expression [119]; and LEF1, EBF1, CDKN2A, FBXW7, CDKN2A/2B, IKZF1, PAX5, RUNX1, and ETV6 deletions [2]. ZNF384r should be detected using WTS or DNA-seq while taking into account that specific genomic profiles could benefit from targeted therapies such as histone deacetylase or JAK-STAT pathway inhibitors [120,121].

4.3.6. MEF2D-Rearranged (1q22)

The incidence ranges from 2% to 5% of all B-ALL cases, and it is more frequent in children and AYAs. The immunophenotype is characterized by CD10-negative status, CD38-positive status, and, often, aberrant CD5 expression [122,123]. MEF2D rearrangements preserve the MEF2D DNA-binding domain, resulting in enhanced transcriptional activity, including that of 3′ partner genes such as BCL9, HNRNPUL1, HNRNPM, DAZAP1, FOXJ2, HNRNPH1, STAT6, SF1R, SS18, and CSF1R [124,125]. Those fusions are usually mutually exclusive with the known risk stratification of chromosomal translocations (at least at the cytogenetic level), except for MEF2D::CSF1R, which display a Ph-like profile [78]. Additional genetic alterations include copy number oscillations involving multiple chromosomes, suggesting chromothripsis, IKZF1 and CDKN2A/B CNVs, dysregulation of RNA stability, NRAS mutations [123,124,126], or GATA3 and HDAC9 overexpression. This suggests that these patients could benefit from histone deacetylase inhibitors such as Panobinostat [2,16,123]. The clinical features are still not well defined. Nevertheless, it has been considered as a high-risk molecular subtype, given that patients may have elevated white blood cells, ineffective stem cell transplantation, and relapse rates around 53.3% [125,127,128]. Although FISH probes exist for the detection of fusions such as MEF2D::BCL9, other techniques, such as WGS, WTS, or OGM, could be used to identify other less frequent or novel rearrangements.

4.3.7. NUTM1-Rearranged (15q14)

This has been reported in about 3% of B-ALL cases, and it is more common in infants (<1 year) than in children (0.4–0.9%), while it is rare in adults [92,118], and it is associated with a good prognosis. According to Judith M. Boer et al. (2021), the 4-year event-free and overall survival rates were 92.1% and 100%, respectively, and 91.1% of patients had an MR of <0.05% [129]. This rearrangement might involve many 5′ partners, including BRD9, BRD4, IKZF1, CUX1, AFF1, ZNF618, ACIN1, and SLC12A6, increased aberrant NUTM1 expression, and altered global chromatin acetylation. This fact might confer histone deacetylase inhibitors with sensitivity [129,130,131]. The downstream effects of NUTM1r include the down-regulation of the NOTCH and Hedgehog signaling pathways, proto-oncogene BMI1 up-regulation, and high expression of clustered genes on chromosome band 10p12.31-12.2, including HOXA gene cluster. In fact, in 2021, two biological subgroups within the NUTM1r subtype were suggested: (1) HOXA9-positive, with a limited number of partners (ACIN1, CUX1, BRD9, and AFF1), which is more prevalent in infants of <9 months of age and was reported in association with CRLF2r; (2) HOXA9-negative, which occurs in infants approaching 1 year of age. Almost half of all NUTM1r pediatric cases occur in this subgroup. Until now, no other somatic sentinel alterations have been reported, suggesting that NUTM1r could be a possible oncogenic driver event [131,132,133].

4.3.8. UBTF::ATXN7L3/PAN3, CDX2 (Also Known as “CDX2/UBTF”)

This entity has been reported in approximately 1% of ALL cases and is more common in AYAs and females. It is characterized by a CD10-negative and IgM-positive immunophenotype. This genotype is featured by two co-occurring genetic lesions: a cryptic deletion at the 17q21.31 locus involving loss of the last four UBTF exons (exon 18 to exon 21) and an intergenic sequence, generating a chimeric UBTF::ATXN7L3 protein [134]; a deletion of approximately 41Kb (common and minimal-size deletion) encompassing the intergenic region between FLT3 and PAN3, resulting in monoallelic ectopic expression of CDX2 [135]. Nevertheless, it must be analyzed rigorously that CDX2 overexpression and other 13q12.2 microdeletions that deregulate FLT3 have been reported in B-ALL [136]. Moreover, individual analysis of CDX2 has limited applicability for the identification of this molecular subtype, but CDX2 overexpression has been reported in other leukemias associated with aberrant HOX gene expression [137,138,139]. CDX2/UBTF cases may harbor additional genetic alterations, including 1q gain and CXCR4-activating mutations; PAX5, VPREB1, and/or IKZF1 deletions; and PRY2 and CD9 overexpression. Furthermore, a ChiP-seq analysis showed that 13q12 region deletion alters DNA topology, and that the epigenetic pattern can be altered, as H3K4me3 and H3K27ac spikes have been observed in the CDX2, FLT3, and PAN3 promoters [135]. Regarding risk, this subtype is classified as having poor prognosis. These patients have a higher risk of failing the first induction course and postinduction MRD, resulting in a significantly higher cumulative relapse incidence [134].

4.4. Point Mutations

4.4.1. PAX5P80R

PAX5 mutations used to be considered as secondary events in the leukemogenesis process. However, the identification of a group with PAX5 mutation (P80R) without associations with any other typical features, and a different expression profile according to WES, allowed for the establishment of this new ALL subtype [37]. P80R is the most common PAX5 mutation; it is observed in around 2–5% of B-ALL cases and is more frequent in adults [140,141]. Some cases may present biallelic inactivation of PAX5 due to wild-type allele deletion, homozygous CDKN2A/B or IKZF1 deletion, and hotspot-activating mutations in RAS-signaling genes. GEP studies also reveal a JAK-STAT-signaling alteration [37,142,143]. As with patients with the DUX4r and ZNF384r subtypes, these patients frequently undergo a monocytic switch with a CD2-positive, CD20-low, CD45-high, CD13-negative, CD33-positive, and CD10-high profile [36,83,141]. The prognostic value is variable according to age. Some authors consider it to be an intermediate risk in children and to have a good prognosis for adults [141,142].

4.4.2. IKZF1N159Y

This represents approximately 1% of ALL cases and is characterized by a missense mutation that affects the DNA-binding domain [36,118]. In general, somatic IKZF1 alterations are associated with high-risk B-ALL with poor response to therapy, but the prognostic impact is still controversial [144,145]. IKZF1N159Y is associated with gain of chromosome 21, KRAS mutations, B-cell receptor downregulation, JAK-STAT-signaling alteration (FLT3, FLT4, and STAT5A) [118], up-regulation of the IKZF1-interacting gene YAP1 (a transcriptional co-activator that can drive KRAS-induced transformation by rescuing cell viability in KRAS-dependent cells), ARHGEF28, and overexpression of the chromatin remodeler SALL1 over-expression [36,84]. Previous studies have reported that SALL1 can recruit histone deacetylase (HDAC) to mediate transcriptional repression. Its promoter is frequently methylated in BCP ALL [35]. On the other hand, partial IKZF1 partial tandem duplications (IKZF1-PTD) ranging in size from 344 bp to 396 bp and span the N159Y variant have been observed in exon 5 [146].

4.5. Other Emerging B-ALL Subtypes

4.5.1. PAX5-Altered

PAX5 is considered a master regulator of B cell development and maintenance and is the most commonly altered gene in ALL, as it is detected in approximately one-third of cases [147]. It is associated with poor prognosis, but other clinical variables must be taken into consideration [142]. This molecular subtype includes CNVs (whole-gene, partial, or intragenic regions), translocations, different point mutations to P80R that generate haploinsufficiency, and combinations thereof that may result in biallelic inactivation [143].
Monoallelic deletions of PAX5 (focal deletion, loss of gene regions of 9p, or chr9 monosomy) are the most common PAX5 variants in B-ALL and are considered a secondary or late event. This has been reported in more than 50% of BCR::ABL1 cases, about 20% of TCF3::PBX1 patients, and as a co-lesion in Ph-like [141]. It is often associated with complete CDKN2A/B loss; STAT5 activation; IL-7 pathway hyperactivation; RB1, BTG1, IKZF1, VPREB1, and BTLA deletions; and/or epigenetic alterations (e.g., KDM6A, KMT2A) [142]. PAX5 fusions are less common (2%-3% of B-ALL cases) but usually appear in cases with normal karyotypes; therefore, they are considered a founding effect in the process of leukemogenesis. The majority of PAX5 rearrangements result in chimeric genes that retain the DNA-binding paired box domain and nuclear localization domain. PAX5 3′ partner genes include ETV6, JAK2, ZCCHC7, NOL4L, and AUTS2, in addition to, less commonly, ZNF521, FOXP1, DACH1, NCOA5, ELN, POM121, PML, MLLT3, CHFR, SOX5, and POM121C. As a result, these fusion proteins may act as dominant negative inhibitor [148]. PAX5 alterations also include intragenic amplifications (PAX5AMP) and germline mutations. The incidence PAX5-ITD is around 0.5–1.4% in B-ALL and consist in an approximately 4–5 copies in PAX5 exons 2 to 5. The functional role of PAX5-iAmp should be further investigated [149,150]. The somatic variants include G183S, R38H, and R140L, as they have been shown to predispose one to familial ALL-B and to cause, at the functional level, loss of B-cell subtypes and immune deficiency. In fact, the ICC considers that germline alterations in PAX5 constitute a distinct entity [140].

4.5.2. Non-Ph-like CRLF2r and “Double-Hit” BCL2/MYC (IGH/BCL2 and 8q24/MYC Rearrangement)

This genotype occurs in approximately 5% of B-ALL cases and is associated with alterations in the JAK/STAT or PI3K/mTOR pathways, high relapse frequency, and, therefore, poorer prognosis [64]. BCL2/MYC is a rare event, but the characterization of MYCr is relevant, given that the fusion partner gene could define the role of MYC and its level of expression during leukemogenesis, increasing early-relapse risk and causing a poor prognosis [89,151].

5. New Insights into Genomic Profiling of T-ALL

Establishing molecular subtypes in T-lineage acute leukemia is more complex. Based on immunophenotypic criteria, a “gray zone” is found, given that some cases do not have clear lineage-specific antigens [152]. Initially, four main T-ALL subtypes were proposed: mature (sCD3 positive), cortical (CD1a+), pre-T (cCD3+ plus CD2+, CD5+, and/or CD8+), and pro-T ALL (cCD3+ and CD7+); however, with the integration with RNA-microarrays, the T-ALL subtypes began to be better characterized [153,154,155].
The most common alterations in T-ALL includes the TAL gene family (TAL1, TAL2, and LYL1 proto-oncogenes) and LMO family rearrangements [156] (Figure 3). In children, these alterations account for approximately 30%, and in adults, the incidence is lower (~12%). These rearrangements may be accompanied by CASP8AP2 (20%) or PHF6 mutations and PTEN deletions. Other common alterations include HOXA (HOXA10, HOXA9) rearrangements or overexpression (20–30% of cases) [157,158], accompanied by KMT2A, MYB, and SP1 rearrangements; PICALM::MLLT10 or SET::NUMP214 fusion; or ETV6, CNOT3, EZH2, JAK3, and STAT5B mutations [159,160]. The mutational T-ALL landscape also includes CDKN2AB, TLX1, NKX1, and KMT2A deletions; ABL1 rearrangements; and IL7R, JAK3, JAK1, WT1, FBXW7 [99], PTEN, RPL10 [155], or NOTCH (9q34.3) mutations [161,162]. This variant is associated with CDKN2A and FBXW7 deletions and occurs in approximately 50% to 60% of pediatric T-ALL cases. These deletions are associated with poorer prognosis [163]. TLX3 rearrangements are also common in T-ALL. However, more studies are needed to compare the clinical and prognostic implications, as the results are inconsistent [164].
According to WHO-HAEM5, there are only two T-ALL groups that can be clearly differentiated. T-ALL NOS, and T-cell precursor early acute lymphoblastic leukemia (ETP-ALL) are characterized by an immunophenotype with negative or absent CD1a and CD8 (<5% positive cells), negative or low CD5 (<75% blast cells), and expression of one or more myeloid (CD11b, CD13, CD33, CD117) or stem cell markers (CD34, HLA-DR) [155,165,166]. The genomic and transcriptomic profiles are characterized by activating mutations in genes regulating cytokine receptor signaling (RAS BRAF, NRAS, KRAS, FLT3, JAK1, JAK3, and SH2B3) (~67% of cases), inactivating lesions that affect hematopoietic development (GATA3, ETV6, RUNX1, IKZF1, and EP300), or histone-modifying genes (EZH2, EED, SUZ12, SETD2 and EP300), as well as MLH3, SUZ12, PHF6, WT1, DNM2, ECT2L, or REL mutations [159,160,166]. Other relevant mutations in genes such as RELN, MIR1297, and IL7R were proposed by St Jude Children’s Research Hospital in the Cancer Genome Project by performing WGS for matched leukemic and normal DNA from pediatric ETP-ALL patients. Two hotspots of missense and in-frame insertion–deletion mutations (IL241-242 and VA253-254) of IL7R were reported [167].

6. Conclusions

Recent advances in NGS and their translation to clinical practice have improved the characterization of the ALL patient mutational landscape, risk stratification, and management. This review highlights new and emerging subtypes of ALL included in recent system classifications, such as WHO-HAEM5 (2022) and ICC (2022), and their prognostic value based on their molecular features, including somatic mutations, gene fusion, epigenetic markers, transcriptional profiles, and alterations in molecular pathways that allow for defining phenocopies.
NGS technologies, such as whole-genome sequencing, targeted sequencing, transcriptome analysis, chromatin immunoprecipitation followed by sequencing (ChIP-seq), ATAC-seq, among others, have proven to be powerful tools for advancing the knowledge of ALL. In the research field, these technologies have allowed for genomic characterization even at the single-cells level to identify driver mutations related to disease development in high-risk subtypes such as KMT2Ar [168,169], or genetic alterations relevant to disease progression and treatment response in genes such as TTN, FLT3, TP53, MUC16, and EPPK1 in B-ALL, and NOTCH1, FBXW7, TTN, MUC16, and PHF6 in T-ALL [170]. In this sense, the prognostic value of NGS in ALL also lies in its potential to improve patient characterization at diagnosis and provide insight into the genetic heterogeneity of the disease, including the identification of alterations in genes such as IKZF1, TP53, CRLF2, and JAK2, which are associated with poor prognosis in B-ALL, to detect differential gene expression profiles or epigenetic patterns that may modulate leukemogenesis and treatment response, such as NOTCH1 in T-ALL, to establish clonal heterogeneity and trace evolutionary trajectories that are useful for identifying alterations associated with relapse risk [171,172], as well as to characterize of differential transcriptomic profiles to identify novel risk subgroups.
Due to their impact, some NGS applications have been transferred to the clinical setting to evaluate the minimal residual disease and identify high-risk variants or subtypes with greater precision, and they have been incorporated into guidelines or recommendations, such as those of the European Leukemia Network-ELN (2022), and clinical trials, such as the PETHEMA protocol (ALL-2019), SEHOP/PETHEMA, and other protocols evaluating responses to immunotherapy. Clinical services in referral hospitals have adopted the use of technologies due to their ability to improve diagnostic accuracy, risk stratification, and minimal residual disease (MRD) as a critical factor in disease monitoring due to their higher sensitivity (factor 1 × 10−6) compared to polymerase chain reaction (PCR) or flow cytometry assays. In clinical practice, the use of the NGS-based ClonoSEQ assay has improved the MDR detection, even using peripheral blood samples. This assay was the first NGS-based test approved by the FDA for ALL, allowing for deep response measurement, early relapse detection, and broader genetic analysis to identify emerging mutations that may contribute to treatment resistance [173]. In addition, it is used for pre- and post-transplant monitoring in high-risk patients. This application has been included in some recommendations, such as NCCN (2024) and European LeukemiaNet recommendations-ELN (2022) [22,29]. However, other guidelines, such as the ESMO Clinical Practice Guideline (2024), highlight some considerations on the MDR detection by NGS among high-risk genetic subtypes such as Ph+, suggesting that real-time quantitative PCR for BCR::ABL and PCR for Ig/TR rearrangements are better, and NGS is useful for the delineation of CD19 clones prior to CAR-T therapy and response assessment to these therapies [174]. In addition, the prognostic value of MDR detection by NGS is also being evaluated in clinical trials, with different approaches, including evaluating the efficacy and safety of treatment schemes such as combination chemotherapy with CAR-T cells in B-ALL patients (NCT06481241), testing and monitoring B-Cell recovery to guide management after Chimeric Antigen Receptor T (CART) cell administration, and monitoring B-ALL relapse in children and young adults with B-ALL (NCT05621291).
NGS techniques are also being used in clinical trials for targeted and personalized treatments. The identification of specific mutations by NGS allows the use of targeted therapies in patients with mutations in ABL1 or JAK2, improving response rates and reducing side effects compared to conventional chemotherapy. In this sense, NGS is useful for treatment decision making and contributes to prognosis. Other applications include the identification of phenocopies as Ph-like in B-ALL and their prognostic assessment (NCT02580981), the assessment of chimerism and relapse post Bone Marrow/Hematopoietic cell transplant (HCT) (NCT04635384), which analyzes SNP loci to quantify donor and recipient cells by measuring genomic DNA prior to transplantation, and the characterization of the mutational landscape or transcriptional profiles in B-ALL, such as the genetic study of familial Acute Lymphoblastic Leukemia (NCT03067584) to identify gene variants that are inherited and influence the risk of developing acute lymphoblastic leukemia, including novel SNPs as a risk. In T-ALL, clinical trials are also underway to evaluate the safety and efficacy of multiple base-edited allogeneic Anti-CD7 CAR-T cells in relapsed/refractory patients (NCT05885464), or to assess the effect of treatment as a prognostic HDACi Chidamide targeted therapy for adult T-LBL/ALL (NCT03564704).
Another concrete application in clinical practice is multicenter studies aimed at standardizing NGS protocols for the management of ALL patients at diagnosis and treatment. The inclusion of targeted panel sequencing in protocols, such as those of the PETHEMA group, and, in particular, the assessment of TP53 status in B-ALL, and NOTCH1, KRAS, NRAS, and FBXW7 in T-ALL (NCT04179929), has made it possible to evaluate mutational profiles and make decisions for the benefit of patients. Hospitals and research institutions have developed risk classifiers based on NGS data that allow clinicians to tailor treatments based on each patient’s molecular profile, an increasingly relevant approach as precision medicine advances in hematologic cancers. Advances in bioinformatics have also enabled the development of algorithms such as ALLsort based on RNAseq data that allow better classification of ALL patients [82]. In addition, the technical and clinical validation of TGS to improve the molecular characterization of ALL patients based on mutation profiles and CNVs. Others panel such as Archer™ FUSIONPlex™ Leukemia panel also allows for the identification of fusions and gene expression profiling, as well as splicing, exon-skipping, SNVs, indels, and others.
Currently, NGS methods are playing an increasingly important role in ALL both in genotyping and prognostic evaluation. While NGS effectively detects a broad range of genetic alterations, it remains complementary to traditional cytogenetic methods due to limitations in identifying certain structural variants and copy number variations (CNVs). The use of NGS technologies with gold-standard techniques is the best alternative for the management of ALL patients in the context of personalized medicine. A combined approach in which NGS and OGM are used alongside established cytogenetic techniques, such as karyotyping, fluorescence in situ hybridization (FISH), and SNP arrays, to enhance diagnostic precision and prognostic assessment. This combination approach is considered the current gold standard in ALL genotyping, providing a comprehensive understanding of the genomic landscape essential for accurate risk stratification and personalized treatment planning.
Finally, all the advances and applications mentioned in this review highlight the important prognostic and practical impact of NGS in the management of ALL patients, making it a useful cornerstone for treatment personalization and suggesting that its incorporation into routine clinical practice can significantly optimize outcomes in patients with ALL in the context of personalized medicine.

Author Contributions

V.R.M.: Conceptualization and achievement of overall research objectives, review process and data collection and synthesis, drafting of initial article and publication. R.B.S.: Conceptualization and objectives, methodology direction and mentoring, management and coordination responsibility, planning and execution, writing—reviewing and editing. J.M.H.R.: Conceptualization and objectives, review and editing, support in the acquisition of funds for the project under which this publication is framed and financed. A.H.-S.: Conceptualization and review; J.N.A. and I.M.M.: Development of the research project under which this publication is framed and funded. Á.V.R.: Figures conceptualization, design, review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Fundación Mutua Madrileña FMM21/002 (AP176752021), Fondos FEDER (EU) and Consejería Educación Junta Castilla y León (SA118P20, SA198P24), Instituto de Investigación Biomédica de Salamanca-IBSAL-Herencia Juan Salvador Escudero (IBPED21/00001), Proyectos de Investigación del SACYL, Spain, Gerencia Regional de Salud de Castilla y León, GRS 2385/A/21, GRS 2386/A/21, GRS2506/A22, GRS2823/A1/2023. Co-funded by the European Union, by grants (RD12/0036/0069) from Red Temática de Investigación Cooperativa en Cáncer (RTICC) and Centro de Investigación Biomédica en Red de Cáncer (CIBERONC CB16/12/00233). AHS is supported by a Río Hortega contract (CM23/00101) from the Instituto de Salud Carlos III (ISCIII). VRM was supported by Colfuturo Scholarship-Credit as a collaborative and cooperative program with the National Government of Colombia, through the Ministry of Science, Technology and Innovation; and Sura Colombia as sponsorship.

Acknowledgments

We thank Sandra Santos, Cristina Miguel, Sara González, Clara Cabrero Tejero, Irene Rodríguez, Mª Ángeles Ramos, Mª Almudena Martín, Ana Díaz, Ana Simón, Filomena Corral, María del Pozo and Sandra Pujante, from the Cancer Research Center of Salamanca, Spain, for their technical support.

Conflicts of Interest

JMHR as BMS: Honoraria, Membership on an entity’s Board of Directors or advisory committees, Other: Research Support, Speakers Bureau; Pfizer: Honoraria, Membership on an entity’s Board of Directors or advisory committees; GSK: Consultancy, Honoraria; Amgen: Honoraria, Membership on an entity’s Board of Directors or advisory committees, Speakers Bureau; Celgene: Honoraria, Membership on an entity’s Board of Directors or advisory committees, Other: Research support, Speakers Bureau; Novartis: Honoraria, Membership on an entity’s Board of Directors or advisory committees, Other: Research Support; Janssen: Membership on an entity’s Board of Directors or advisory committees, Speakers Bureau; Roche: Membership on an entity’s Board of Directors or advisory committees, Speakers Bureau; Abbvie: Membership on an entity’s Board of Directors or advisory committees, Speakers Bureau; AstraZeneca: Membership on an entity’s Board of Directors or advisory committees, Speakers Bureau; Beigene: Membership on an entity’s Board of Directors or advisory committees; Lilly: Membership on an entity’s Board of Directors or advisory committees; Gilead: Membership on an entity’s Board of Directors or advisory committees, Speakers Bureau; Takeda: Membership on an entity’s Board of Directors or advisory committees, Speakers Bureau; Jazz Pharmaceuticals: Membership on an entity’s Board of Directors or advisory committees; Rovi: Membership on an entity’s Board of Directors or advisory committees; Incyte: Membership on an entity’s Board of Directors or advisory committees.

Disclaimer

The authors declare that the access and use of the data complies with the standards of transparency, intellectual property and copyright; and that they agree with the list and contributions of each one of the authors. Furthermore, the project in which this publication is framed has the endorsement of the ethic committee and the approval for publication.

References

  1. Aleem, A.; Haque, A.R.; Roloff, G.W.; Griffiths, E.A. Application of Next-Generation Sequencing-Based Mutational Profiling in Acute Lymphoblastic Leukemia. Curr. Hematol. Malig. Rep. 2021, 16, 394–404. [Google Scholar] [CrossRef] [PubMed]
  2. Lejman, M.; Chałupnik, A.; Chilimoniuk, Z.; Dobosz, M. Genetic Biomarkers and Their Clinical Implications in B-Cell Acute Lymphoblastic Leukemia in Children. Int. J. Mol. Sci. 2022, 23, 2755. [Google Scholar] [CrossRef] [PubMed]
  3. Inaba, H.; Greaves, M.; Mullighan, C.G. Acute lymphoblastic leukaemia. Lancet 2013, 381, 1943–1955. [Google Scholar] [CrossRef] [PubMed]
  4. Malard, F.; Mohty, M. Acute lymphoblastic leukaemia. Lancet 2020, 395, 1146–1162. [Google Scholar] [CrossRef]
  5. Raetz, E.A.; Teachey, D.T. T-cell acute lymphoblastic leukemia. Hematol. Am. Soc. Hematol. Educ. Program. 2016, 1, 580–588. [Google Scholar] [CrossRef]
  6. Terwilliger, T.; Abdul-Hay, M. Acute lymphoblastic leukemia: A comprehensive review and 2017 update. Blood Cancer J. 2017, 7, e577. [Google Scholar] [CrossRef]
  7. Schaaf, J.; Sedlmayr, M.; Schaefer, J.; Storf, H. Diagnosis of Rare Diseases: A scoping review of clinical decision support systems. Orphanet J. Rare Dis. 2020, 15, 263. [Google Scholar] [CrossRef]
  8. Katz, A.J.; Chia, V.M.; Schoonen, W.M.; Kelsh, M.A. Acute lymphoblastic leukemia: An assessment of international incidence, survival, and disease burden. Cancer Causes Control 2015, 26, 1627–1642. [Google Scholar] [CrossRef]
  9. Sasaki, K.; Jabbour, E.; Short, N.J.; Jain, N.; Ravandi, F.; Pui, C.; Kantarjian, H. Acute lymphoblastic leukemia: A population-based study of outcome in the United States based on the surveillance, epidemiology, and end results (SEER) database, 1980–2017. Am. J. Hematol. 2021, 96, 650–658. [Google Scholar] [CrossRef] [PubMed]
  10. Downing, J.R.; Wilson, R.K.; Zhang, J.; Mardis, E.R.; Pui, C.-H.; Ding, L.; Ley, T.J.; Evans, W.E. The pediatric cancer genome project. Nat. Genet. 2012, 44, 619–622. [Google Scholar] [CrossRef]
  11. Hunger, S.P.; Mullighan, C.G. Acute lymphoblastic leukemia in children. N. Engl. J. Med. 2015, 373, 1541–1552. [Google Scholar] [CrossRef] [PubMed]
  12. Beneduce, G.; De Matteo, A.; Stellato, P.; Testi, A.M.; Bertorello, N.; Colombini, A.; Putti, M.C.; Rizzari, C.; Cesaro, S.; Cellini, M.; et al. Blinatumomab in Children and Adolescents with Relapsed/Refractory B Cell Precursor Acute Lymphoblastic Leukemia: A Real-Life Multicenter Retrospective Study in Seven AIEOP (Associazione Italiana di Ematologia e Oncologia Pediatrica) Centers. Cancers 2022, 14, 426. [Google Scholar] [CrossRef] [PubMed]
  13. Abdelmabood, S.; Fouda, A.E.; Boujettif, F.; Mansour, A. Treatment outcomes of children with acute lymphoblastic leukemia in a middle-income developing country: High mortalities, early relapses, and poor survival. J. Pediatr. (Rio J.) 2020, 96, 108–116. [Google Scholar] [CrossRef]
  14. Lennmyr, E.; Karlsson, K.; Ahlberg, L.; Garelius, H.; Hulegårdh, E.; Izarra, A.S.; Joelsson, J.; Kozlowski, P.; Moicean, A.; Tomaszewska-Toporska, B.; et al. Survival in adult acute lymphoblastic leukaemia (ALL): A report from the Swedish ALL Registry. Eur. J. Haematol. 2019, 103, 88–98. [Google Scholar] [CrossRef]
  15. Iacobucci, I.; Mullighan, C.G. Genetic basis of acute lymphoblastic leukemia. J. Clin. Oncol. 2017, 35, 975–983. [Google Scholar] [CrossRef]
  16. Tran, T.H.; Hunger, S.P. The genomic landscape of pediatric acute lymphoblastic leukemia and precision medicine opportunities. Semin. Cancer Biol. 2022, 84, 144–152. [Google Scholar] [CrossRef] [PubMed]
  17. Akkari, Y.M.N.; Baughn, L.B.; Dubuc, A.M.; Smith, A.C.; Mallo, M. Guiding the global evolution of cytogenetic testing for hematologic malignancies. Blood 2020, 139, 2273–2284. [Google Scholar] [CrossRef]
  18. Moorman, A.V.; Harrison, C.J.; Buck, G.A.N.; Richards, S.M.; Secker-Walker, L.M.; Martineau, M.; Vance, G.H.; Cherry, A.M.; Higgins, R.R.; Fielding, A.K.; et al. Karyotype is an independent prognostic factor in adult acute lymphoblastic leukemia (ALL): Analysis of cytogenetic data from patients treated on the Medical Research Council (MRC) UKALLXII/Eastern Cooperative Oncology Group (ECOG) 2993 trial. Blood 2007, 109, 3189–3197. [Google Scholar] [CrossRef] [PubMed]
  19. Bendari, M.; Sraidi, S.; Khoubila, N. Genetic Abnormalities in ALL. In Cytogenetics-Classical and Molecular Strategies for Analysing Heredity Material. Cap 8; IntechOpen: London, UK, 2021. [Google Scholar] [CrossRef]
  20. Othman, M.A.K.; Melo, J.B.; Carreira, I.M.; Rincic, M.; Glaser, A.; Grygalewicz, B.; Gruhn, B.; Wilhelm, K.; Rittscher, K.; Meyer, B.; et al. High rates of submicroscopic aberrations in karyotypically normal acute lymphoblastic leukemia. Mol. Cytogenet. 2015, 8, 45. [Google Scholar] [CrossRef]
  21. Hunger, S.P.; Mullighan, C.G. Redefining ALL classification: Toward detecting high-risk ALL and implementing precision medicine. Blood 2015, 125, 3977–3987. [Google Scholar] [CrossRef]
  22. Brown, P.A.; Shah, B.; Advani, A.; Aoun, P.; Boyer, M.W.; Burke, P.W.; DeAngelo, D.J.; Dinner, S.; Fathi, A.T.; Gauthier, J.; et al. Acute lymphoblastic leukemia, version 2.2021. J. Natl. Compr. Cancer Netw. 2021, 19, 1079–1109. [Google Scholar] [CrossRef] [PubMed]
  23. Neumann, M.; Vosberg, S.; Schlee, C.; Heesch, S.; Schwartz, S.; Gökbuget, N.; Hoelzer, D.; Graf, A.; Krebs, S.; Bartram, I.; et al. Mutational spectrum of adult T-ALL. Oncotarget 2014, 6, 2754–2766. [Google Scholar] [CrossRef] [PubMed]
  24. Zhang, H.-H.; Wang, H.-S.; Qian, X.-W.; Zhu, X.-H.; Miao, H.; Yu, Y.; Meng, J.-H.; Le, J.; Jiang, J.-Y.; Cao, P.; et al. Ras pathway mutation feature in the same individuals at diagnosis and relapse of childhood acute lymphoblastic leukemia. Transl. Pediatr. 2020, 9, 4–12. [Google Scholar] [CrossRef] [PubMed]
  25. Pavlovic, S.; Kotur, N.; Stankovic, B.; Zukic, B.; Gasic, V.; Dokmanovic, L. Pharmacogenomic and pharmacotranscriptomic profiling of childhood acute lymphoblastic leukemia: Paving the way to personalized treatment. Genes 2019, 10, 191. [Google Scholar] [CrossRef] [PubMed]
  26. Mei, L.; Ontiveros, E.P.; Griffiths, E.A.; Thompson, J.E.; Wang, E.S.; Wetzler, M. Pharmacogenetics predictive of response and toxicity in acute lymphoblastic leukemia therapy. Blood Rev. 2015, 29, 243–249. [Google Scholar] [CrossRef]
  27. Nordlund, J.; Syvänen, A.C. Epigenetics in pediatric acute lymphoblastic leukemia. Semin. Cancer Biol. 2018, 51, 129–138. [Google Scholar] [CrossRef]
  28. Drożak, P.; Bryliński, Ł.; Zawitkowska, J. A Comprehensive Overview of Recent Advances in Epigenetics in Pediatric Acute Lymphoblastic Leukemia. Cancers 2022, 14, 5384. [Google Scholar] [CrossRef]
  29. Gökbuget, N.; Boissel, N.; Chiaretti, S.; Dombret, H.; Doubek, M.; Fielding, A.K.; Foà, R.; Giebel, S.; Hoelzer, D.; Hunault, M.; et al. Management of ALL in adults: 2024 ELN recommendations from a European expert panel. Blood 2024, 143, 1903–1930. [Google Scholar] [CrossRef]
  30. Schwab, C.; Cranston, R.E.; Ryan, S.L.; Butler, E.; Winterman, E.; Hawking, Z.; Bashton, M.; Enshaei, A.; Russell, L.J.; Kingsbury, Z.; et al. Integrative genomic analysis of childhood acute lymphoblastic leukaemia lacking a genetic biomarker in the UKALL2003 clinical trial. Leukemia 2023, 37, 529–538. [Google Scholar] [CrossRef]
  31. Wu, C.; Li, W. Genomics and pharmacogenomics of pediatric acute lymphoblastic leukemia. Crit. Rev. Oncol. Hematol. 2018, 126, 100–111. [Google Scholar] [CrossRef]
  32. Abaji, R.; Krajinovic, M. Pharmacogenetics of asparaginase in acute lymphoblastic leukemia. Cancer Drug Resist. 2019, 2, 242–255. [Google Scholar] [CrossRef] [PubMed]
  33. Jabbour, E.; Short, N.J.; Jain, N.; Haddad, F.G.; Welch, M.A.; Ravandi, F.; Kantarjian, H. The evolution of acute lymphoblastic leukemia research and therapy at MD Anderson over four decades. J. Hematol. Oncol. 2023, 16, 22. [Google Scholar] [CrossRef]
  34. Pui, C.H.; Evans, W.E. A 50-year journey to cure childhood acute lymphoblastic leukemia. Semin. Hematol. 2013, 50, 185–196. [Google Scholar] [CrossRef] [PubMed]
  35. Inaba, H.; Azzato, E.M.; Mullighan, C.G. Integration of next-generation sequencing to treat acute lymphoblastic leukemia with targetable lesions: The St. Jude Children’s Research Hospital approach. Front. Pediatr. 2017, 5, 258. [Google Scholar] [CrossRef] [PubMed]
  36. Li, J.-F.; Dai, Y.-T.; Lilljebjörn, H.; Shen, S.-H.; Cui, B.-W.; Bai, L.; Liu, Y.-F.; Qian, M.-X.; Kubota, Y.; Kiyoi, H.; et al. Transcriptional landscape of B cell precursor acute lymphoblastic leukemia based on an international study of 1223 cases. Proc. Natl. Acad. Sci. USA 2018, 115, E11711–E11720. [Google Scholar] [CrossRef]
  37. Passet, M.; Boissel, N.; Sigaux, F.; Saillard, C.; Bargetzi, M.; Ba, I.; Thomas, X.; Graux, C.; Chalandon, Y.; Leguay, T.; et al. PAX5 P80R mutation identifies a novel subtype of B-cellprecursor acute lymphoblastic leukemia with favorable outcome. Blood 2019, 133, 280–284. [Google Scholar] [CrossRef]
  38. Ishida, H.; Iguchi, A.; Aoe, M.; Takahashi, T.; Tamefusa, K.; Kanamitsu, K.; Fujiwara, K.; Washio, K.; Matsubara, T.; Tsukahara, H.; et al. Panel-based next-generation sequencing identifies prognostic and actionable genes in childhood acute lymphoblastic leukemia and is suitable for clinical sequencing. Ann. Hematol. 2019, 98, 657–668. [Google Scholar] [CrossRef]
  39. James, A.R.; Schroeder, M.P.; Neumann, M.; Bastian, L.; Eckert, C.; Gökbuget, N.; Tanchez, J.O.; Schlee, C.; Isaakidis, K.; Schwartz, S.; et al. Long non-coding RNAs defining major subtypes of B cell precursor acute lymphoblastic leukemia. J. Hematol. Oncol. 2019, 12, 8. [Google Scholar] [CrossRef]
  40. Fleur-Lominy, S.S.; Evensen, N.A.; Bhatla, T.; Sethia, G.; Narang, S.; Choi, J.H.; Ma, X.; Yang, J.J.; Kelly, S.; Raetz, E.; et al. Evolution of the epigenetic landscape in childhood B Acute lymphoblastic leukemia and its role in drug resistance. Cancer Res. 2020, 80, 5189–5202. [Google Scholar] [CrossRef]
  41. Erarslan-Uysal, B.; Kunz, J.B.; Rausch, T.; Richter-Pechańska, P.; van Belzen, I.A.; Frismantas, V.; Bornhauser, B.; Ordoñez-Rueada, D.; Paulsen, M.; Benes, V.; et al. Chromatin accessibility landscape of pediatric T-lymphoblastic leukemia and human T-cell precursors. EMBO Mol. Med. 2020, 12, e12104. [Google Scholar] [CrossRef]
  42. Navrkalova, V.; Plevova, K.; Hynst, J.; Pal, K.; Mareckova, A.; Reigl, T.; Jelinkova, H.; Vrzalova, Z.; Stranska, K.; Pavlova, S.; et al. LYmphoid NeXt-Generation Sequencing (LYNX) Panel: A Comprehensive Capture-Based Sequencing Tool for the Analysis of Prognostic and Predictive Markers in Lymphoid Malignancies. J. Mol. Diagn. 2021, 23, 959–974. [Google Scholar] [CrossRef] [PubMed]
  43. Hrabovsky, S.; Vrzalova, Z.; Stika, J.; Jelinkova, H.; Jarosova, M.; Navrkalova, V.; Martenek, J.; Folber, F.; Salek, C.; Horacek, J.M.; et al. Genomic landscape of B-other acute lymphoblastic leukemia in an adult retrospective cohort with a focus on BCR-ABL1-like subtype. Acta Oncol. 2021, 60, 760–770. [Google Scholar] [CrossRef] [PubMed]
  44. Alsuwaidi, L.; Hachim, M.; Senok, A. Novel Markers in Pediatric Acute Lymphoid Leukemia: The Role of ADAM6 in B Cell Leukemia. Front. Cell Dev. Biol. 2021, 9, 706129. [Google Scholar] [CrossRef] [PubMed]
  45. Stewart, J.P.; Gazdova, J.; Darzenta, N.; Wren, D.; Proszek, P.; Fazio, G.; Songia, S.; Alcoceba, M.; Sarasquete, M.E.; Villarese, P.; et al. Validation of the EuroClonality-NGS DNA capture panel as an integrated genomic tool for lymphoproliferative disorders. Blood Adv. 2021, 5, 3188–3198. [Google Scholar] [CrossRef]
  46. van der Velden, V.H.; Brüggemann, M.; Cazzaniga, G.; Scheijen, B.; Tops, B.; Trka, J.; Pal, K.; Hänzelmann, S.; Fazio, G.; Songia, S.; et al. Potential and pitfalls of whole transcriptome-based immunogenetic marker identification in acute lymphoblastic leukemia; a EuroMRD and EuroClonality-NGS Working Group study. Leukemia 2021, 35, 924–928. [Google Scholar] [CrossRef]
  47. Hetzel, S.; Mattei, A.L.; Kretzmer, H.; Qu, C.; Chen, X.; Fan, Y.; Wu, G.; Roberts, K.G.; Luger, S.; Litzow, M.; et al. Acute lymphoblastic leukemia displays a distinct highly methylated genome. Nat. Cancer 2022, 3, 768–782. [Google Scholar] [CrossRef]
  48. Chaber, R.; Gurgul, A.; Tabarkiewicz, J.; Wróbel, G.; Szmatoła, T.; Jasielczuk, I.; Haus, O.; Lejman, M.; Rybka, B.; Ryczan-Krawczyk, R.; et al. MicroRNA gene methylation landscape in pediatric B-cell precursor acute lymphoblastic leukemia. Adv. Clin. Exp. Med. 2022, 31, 293–305. [Google Scholar] [CrossRef]
  49. Longjohn, M.N.; Squires, W.R.B.; Christian, S.L. Meta-analysis of microRNA profiling data does not reveal a consensus signature for B cell acute lymphoblastic leukemia. Gene 2022, 821, 146211. [Google Scholar] [CrossRef]
  50. Barnett, K.R.; Mobley, R.J.; Diedrich, J.D.; Bergeron, B.P.; Bhattarai, K.R.; Yang, W.; Crews, K.R.; Manring, C.S.; Jabbour, E.; Paietta, E.; et al. Epigenomic mapping in B-cell acute lymphoblastic leukemia identifies transcriptional regulators and noncoding variants promoting distinct chromatin architectures. bioRxiv 2023, 3, 1–28. [Google Scholar] [CrossRef]
  51. Levy, B.; Kanagal-Shamanna, R.; Sahajpal, N.S.; Neveling, K.; Rack, K.; Dewaele, B.; Weghuis, D.O.; Stevens-Kroef, M.; Puiggros, A.; Mallo, M.; et al. A framework for the clinical implementation of optical genome mapping in hematologic malignancies. Am. J. Hematol. 2024, 99, 642–661. [Google Scholar] [CrossRef]
  52. Sahajpal, N.S.; Mondal, A.K.; Tvrdik, T.; Hauenstein, J.; Shi, H.; Deeb, K.K.; Saxe, D.; Hastie, A.R.; Chaubey, A.; Savage, N.M.; et al. Clinical Validation and Diagnostic Utility of Optical Genome Mapping for Enhanced Cytogenomic Analysis of Hematological Neoplasms. J. Mol. Diagn. 2022, 24, 1279–1291. [Google Scholar] [CrossRef]
  53. Gordeeva, V.; Sharova, E.; Arapidi, G. Progress in Methods for Copy Number Variation Profiling. Int. J. Mol. Sci. 2022, 23, 2143. [Google Scholar] [CrossRef] [PubMed]
  54. Song, Y.; Fang, Q.; Mi, Y. Prognostic significance of copy number variation in B-cell acute lymphoblastic leukemia. Front. Oncol. 2022, 12, 981036. [Google Scholar] [CrossRef] [PubMed]
  55. Alaggio, R.; Amador, C.; Anagnostopoulos, I.; Attygalle, A.D.; Araujo, I.B.D.O.; Berti, E.; Bhagat, G.; Borges, A.M.; Boyer, D.; Calaminici, M.; et al. The 5th edition of the World Health Organization Classification of Haematolymphoid Tumours: Lymphoid Neoplasms. Leukemia 2022, 36, 1720–1748. [Google Scholar] [CrossRef] [PubMed]
  56. Arber, D.A.; Orazi, A.; Hasserjian, R.P.; Borowitz, M.J.; Calvo, K.R.; Kvasnicka, H.M.; Wang, S.A.; Bagg, A.; Barbui, T.; Branford, S.; et al. International Consensus Classification of Myeloid Neoplasms and Acute Leukemias: Integrating morphologic, clinical, and genomic data. Blood 2022, 140, 1200–1228. [Google Scholar] [CrossRef]
  57. MLL Münchner Leukämielabor GmbH. MLL ALL Classification. 2022. Available online: https://www.mll.com/en/acute-lymphoblastic-leukemia-all/acute-lymphoblastic-leukemia-all#:~:text=ALL%3A%20Classification,lymphoblastic%20lymphoma%20(Swerdlow%20et%20al (accessed on 18 March 2024).
  58. Duffield, A.S.; Mullighan, C.G.; Borowitz, M.J. International Consensus Classification of acute lymphoblastic leukemia/lymphoma. Virchows Archiv. 2023, 482, 11–26. [Google Scholar] [CrossRef]
  59. Nishiwaki, S.; Kim, J.H.; Ito, M.; Maeda, M.; Okuno, Y.; Koyama, D.; Ozawa, Y.; Gunji, M.; Osaki, M.; Kitamura, K.; et al. Multi-Lineage BCR-ABL Expression in Philadelphia Chromosome-Positive Acute Lymphoblastic Leukemia Is Associated With Improved Prognosis but No Specific Molecular Features. Front. Oncol. 2020, 10, 586567. [Google Scholar] [CrossRef]
  60. Rathana, K.; Clappier, E. Uncovering new layers of Ph+ ALL biology. Blood 2024, 143, 1322–1323. [Google Scholar] [CrossRef]
  61. Kim, J.C.; Chan-Seng-Yue, M.; Ge, S.; Zeng, A.G.X.; Ng, K.; Gan, O.I.; Garcia-Prat, L.; Flores-Figueroa, E.; Woo, T.; Zhang, A.X.W.; et al. Transcriptomic classes of BCR-ABL1 lymphoblastic leukemia. Nat. Genet. 2023, 55, 1186–1197. [Google Scholar] [CrossRef]
  62. Fedullo, A.L.; Messina, M.; Elia, L.; Piciocchi, A.; Gianfelici, V.; Lauretti, A.; Soddu, S.; Puzzolo, M.C.; Minotti, C.; Ferrara, F.; et al. Prognostic implications of additional genomic lesions in adult Philadelphia chromosomepositive acute lymphoblastic leukemia. Haematologica 2019, 104, 312–318. [Google Scholar] [CrossRef]
  63. Bastian, L.; Beder, T.; Barz, M.J.; Bendig, S.; Bartsch, L.; Walter, W.; Wolgast, N.; Brändl, B.; Rohrandt, C.; Hansen, B.-T.; et al. Developmental trajectories and cooperating genomic events define molecular subtypes of BCR::ABL1-positive ALL. Blood 2023, 143, 1391–1398. [Google Scholar] [CrossRef] [PubMed]
  64. Jain, N.; Roberts, K.G.; Jabbour, E.; Patel, K.; Eterovic, A.K.; Chen, K.; Zweidler-McKay, P.; Lu, X.; Fawcett, G.; Wang, S.A.; et al. Ph-like acute lymphoblastic leukemia: A high-risk subtype in adults. Blood 2017, 129, 572–581. [Google Scholar] [CrossRef] [PubMed]
  65. Tran, T.H.; Hunger, S.P. ABL-class fusion positive acute lymphoblastic leukemia: Can targeting ABL cure ALL? Haematologica 2020, 105, 1754–1757. [Google Scholar] [CrossRef]
  66. Den Boer, M.L.; Cario, G.; Moorman, A.V.; Boer, J.M.; de Groot-Kruseman, H.A.; Fiocco, M.; Escherich, G.; Imamura, T.; Yeoh, A.; Sutton, R.; et al. Outcomes of paediatric patients with B-cell acute lymphocytic leukaemia with ABL-class fusion in the pre-tyrosine-kinase inhibitor era: A multicentre, retrospective, cohort study. Lancet Haematol. 2021, 8, e55–e66. [Google Scholar] [CrossRef]
  67. Roberts, K.G.; Gu, Z.; Payne-Turner, D.; McCastlain, K.; Harvey, R.C.; Chen, I.-M.; Pei, D.; Iacobucci, I.; Valentine, M.; Pounds, S.B.; et al. High Frequency and Poor Outcome of Philadelphia Chromosome-Like Acute Lymphoblastic Leukemia in Adults. J. Clin. Oncol. 2017, 35, 394–401. [Google Scholar] [CrossRef] [PubMed]
  68. Norvilas, R.; Dirse, V.; Semaskeviciene, R.; Mickeviciute, O.; Gineikiene, E.; Stoskus, M.; Vaitkeviciene, G.; Rascon, J.; Griskevicius, L. Low incidence of ABL-class and JAK-STAT signaling pathway alterations in uniformly treated pediatric and adult B-cell acute lymphoblastic leukemia patients using MRD risk-directed approach—A population-based study. BMC Cancer 2021, 21, 326. [Google Scholar] [CrossRef]
  69. Reshmi, S.C.; Harvey, R.C.; Roberts, K.G.; Stonerock, E.; Smith, A.; Jenkins, H.; Chen, I.-M.; Valentine, M.; Liu, Y.; Li, Y.; et al. Targetable kinase gene fusions in high-risk B-ALL: A study from the Children’s Oncology Group. Blood 2017, 129, 3352–3361. [Google Scholar] [CrossRef]
  70. Downes, C.E.; McClure, B.J.; McDougal, D.P.; Heatley, S.L.; Bruning, J.B.; Thomas, D.; Yeung, D.T.; White, D.L. JAK2 Alterations in Acute Lymphoblastic Leukemia: Molecular Insights for Superior Precision Medicine Strategies. Front. Cell Dev. Biol. 2022, 10, 942053. [Google Scholar] [CrossRef]
  71. Cario, G.; Leoni, V.; Conter, V.; Baruchel, A.; Schrappe, M.; Biondi, A. BCR-ABL1-like acute lymphoblastic leukemia in childhood and targeted therapy. Haematologica 2020, 105, 2200–2204. [Google Scholar] [CrossRef]
  72. Płotka, A.; Lewandowski, K. BCR/ABL1-Like Acute Lymphoblastic Leukemia: From Diagnostic Approaches to Molecularly Targeted Therapy. Acta Haematol. 2022, 145, 122–130. [Google Scholar] [CrossRef]
  73. Roberts, K.G.; Janke, L.J.; Zhao, Y.; Seth, A.; Ma, J.; Finkelstein, D.; Smith, S.; Ebata, K.; Tuch, B.B.; Hunger, S.P.; et al. Etv6-ntrk3 induces aggressive acute lymphoblastic leukemia highly sensitive to selective trk inhibition. Blood 2018, 132, 861–865. [Google Scholar] [CrossRef] [PubMed]
  74. Malouf, C.; Ottersbach, K. Molecular processes involved in B cell acute lymphoblastic leukaemia. Cell. Mol. Life Sci. 2018, 75, 417–446. [Google Scholar] [CrossRef] [PubMed]
  75. Chen, D.; Camponeschi, A.; Nordlund, J.; Marincevic-Zuniga, Y.; Abrahamsson, J.; Lönnerholm, G.; Fogelstrand, L.; Mårtensson, I. RAG1 co-expression signature identifies ETV6-RUNX1-like B-cell precursor acute lymphoblastic leukemia in children. Cancer Med. 2021, 10, 3997–4003. [Google Scholar] [CrossRef]
  76. Lilljebjörn, H.; Henningsson, R.; Hyrenius-Wittsten, A.; Olsson, L.; Orsmark-Pietras, C.; Von Palffy, S.; Askmyr, M.; Rissler, M.; Schrappe, M.; Cario, G.; et al. Identification of ETV6-RUNX1-like and DUX4-rearranged subtypes in paediatric B-cell precursor acute lymphoblastic leukaemia. Nat. Commun. 2016, 7, 11790. [Google Scholar] [CrossRef] [PubMed]
  77. Zaliova, M.; Kotrova, M.; Bresolin, S.; Stuchly, J.; Stary, J.; Hrusak, O.; Kronnie, G.T.; Trka, J.; Zuna, J.; Vaskova, M. ETV6/RUNX1-like acute lymphoblastic leukemia: A novel B-cell precursor leukemia subtype associated with the CD27/CD44 immunophenotype. Genes. Chromosomes Cancer 2017, 56, 608–616. [Google Scholar] [CrossRef] [PubMed]
  78. Roberts, K.G. Genetics and prognosis of ALL in chlidren vs adults. Hematol. Am. Soc. Hematol. Educ. Program. 2018, 1, 137–145. [Google Scholar] [CrossRef]
  79. Ryan, S.L.; Peden, J.F.; Kingsbury, Z.; Schwab, C.J.; James, T.; Polonen, P.; Mijuskovic, M.; Becq, J.; Yim, R.; Cranston, R.E.; et al. Whole genome sequencing provides comprehensive genetic testing in childhood B-cell acute lymphoblastic leukaemia. Leukemia 2023, 37, 518–528. [Google Scholar] [CrossRef]
  80. Lee, S.H.R.; Li, Z.; Tai, S.T.; Oh, B.L.Z.; Yeoh, A.E.J. Genetic alterations in childhood acute lymphoblastic leukemia: Interactions with clinical features and treatment response. Cancers 2021, 13, 4068. [Google Scholar] [CrossRef]
  81. Jeha, S.; Choi, J.; Roberts, K.G.; Pei, D.; Coustan-Smith, E.; Inaba, H.; Rubnitz, J.E.; Ribeiro, R.C.; Gruber, T.A.; Raimondi, S.C.; et al. Clinical Significance of Novel Subtypes of Acute Lymphoblastic Leukemia in the Context of Minimal Residual Disease–Directed Therapy. Blood Cancer Discov. 2021, 2, 326–337. [Google Scholar] [CrossRef]
  82. Schmidt, B.M.; Brown, L.M.; Ryland, G.L.; Lonsdale, A.; Kosasih, H.J.; Ludlow, L.E.; Majewski, I.J.; Blombery, P.; Ekert, P.G.; Davidson, N.M.; et al. ALLSorts: An RNA-Seq subtype classifier for B-cell acute lymphoblastic leukemia. Blood Adv. 2022, 6, 4093–4097. [Google Scholar] [CrossRef]
  83. Novakova, M.; Zaliova, M.; Fiser, K.; Vakrmanova, B.; Slamova, L.; Musilova, A.; Brüggemann, M.; Ritgen, M.; Fronkova, E.; Kalina, T.; et al. DUX4r, ZNF384r and PAX5-P80R mutated B-cell precursor acute lymphoblastic leukemia frequently undergo monocytic switch. Haematologica 2021, 106, 2066–2075. [Google Scholar] [CrossRef] [PubMed]
  84. Iacobucci, I.; Kimura, S.; Mullighan, C.G. Biologic and Therapeutic Implications of Genomic Alterations in Acute Lymphoblastic Leukemia. J. Clin. Med. 2021, 10, 3792. [Google Scholar] [CrossRef]
  85. Zaliova, M.; Winkowska, L.; Stuchly, J.; Fiser, K.; Triska, P.; Zwyrtkova, M.; Hrusak, O.; Starkova, J.; Sramkova, L.; Stary, J.; et al. A novel class of ZNF384 aberrations in acute leukemia. Blood Adv. 2021, 5, 4393–4397. [Google Scholar] [CrossRef]
  86. Russell, L.J.; Enshaei, A.; Jones, L.; Erhorn, A.; Masic, D.; Bentley, H.; Laczko, K.S.; Fielding, A.K.; Goldstone, A.H.; Goulden, N.; et al. IGH Translocations Are Prevalent in Teenagers and Young Adults With Acute Lymphoblastic Leukemia and Are Associated With a Poor Outcome. J. Clin. Oncol. 2014, 32, 1453–1462. [Google Scholar] [CrossRef]
  87. Roberts, K.G.; Mullighan, C.G. Genomics in acute lymphoblastic leukaemia: Insights and treatment implications. Nat. Rev. Clin. Oncol. 2015, 12, 344–357. [Google Scholar] [CrossRef] [PubMed]
  88. Parolini, M.; Mecucci, C.; Matteucci, C.; Giussani, U.; Intermesoli, T.; Tosi, M.; Rambaldi, A.; Bassan, R. Highly aggressive T-cell acute lymphoblastic leukemia with t(8;14)(q24;q11): Extensive genetic characterization and achievement of early molecular remission and long-term survival in an adult patient. Blood Cancer J. 2014, 4, e176. [Google Scholar] [CrossRef]
  89. Bomken, S.; Enshaei, A.; Schwalbe, E.C.; Mikulasova, A.; Dai, Y.; Zaka, M.; Fung, K.T.; Bashton, M.; Lim, H.; Jones, L.; et al. Molecular characterization and clinical outcome of B-cell precursor acute lymphoblastic leukemia with IG-MYC rearrangement. Haematologica 2023, 108, 717–731. [Google Scholar] [CrossRef] [PubMed]
  90. Li, Q.; Pan, S.; Xie, T.; Liu, H. MYC in T-cell acute lymphoblastic leukemia: Functional implications and targeted strategi. Blood Sci. 2021, 3, 65–70. [Google Scholar] [CrossRef]
  91. Bhojwani, D.; Yang, J.J.; Pui, C.H. Biology of childhood acute lymphoblastic leukemia. Pediatr. Clin. N. Am. 2015, 62, 47–60. [Google Scholar] [CrossRef]
  92. Schwab, C.J.; Murdy, D.; Butler, E.; Enshaei, A.; Winterman, E.; Cranston, R.E.; Ryan, S.; Barretta, E.; Hawking, Z.; Murray, J.; et al. Genetic characterisation of childhood B-other-acute lymphoblastic leukaemia in UK patients by fluorescence in situ hybridisation and Multiplex Ligation-dependent Probe Amplification. Br. J. Haematol. 2022, 196, 753–763. [Google Scholar] [CrossRef]
  93. Nigro, L.L. Biology of Childhood Acute Lymphoblastic Leukemia. J. Pediatr. Hematol. Oncol. 2013, 35, 245–252. [Google Scholar] [CrossRef] [PubMed]
  94. Brady, S.W.; Roberts, K.G.; Gu, Z.; Shi, L.; Pounds, S.; Pei, D.; Cheng, C.; Dai, Y.; Devidas, M.; Qu, C.; et al. The genomic landscape of pediatric acute lymphoblastic leukemia. Nat. Genet. 2022, 54, 1376–1389. [Google Scholar] [CrossRef] [PubMed]
  95. Herold, T.; Schneider, S.; Metzeler, K.H.; Neumann, M.; Hartmann, L.; Roberts, K.G.; Konstandin, N.P.; Greif, P.A.; Bräundl, K.; Ksienzyk, B.; et al. Adults with philadelphia chromosome–like acute lymphoblastic leukemia frequently have igh-CRLF2 and JAK2 mutations, persistence of minimal residual disease and poor prognosis. Haematologica 2017, 102, 130–138. [Google Scholar] [CrossRef] [PubMed]
  96. Roberts, K.G.; Li, Y.; Payne-Turner, D.; Harvey, R.C.; Yang, Y.-L.; Pei, D.; McCastlain, K.; Ding, L.; Lu, C.; Song, G.; et al. Targetable Kinase-Activating Lesions in Ph-like Acute Lymphoblastic Leukemia. N. Engl. J. Med. 2014, 371, 1005–1015. [Google Scholar] [CrossRef] [PubMed]
  97. Ghazavi, F.; Lammens, T.; Van Roy, N.; Poppe, B.; Speleman, F.; Benoit, Y.; Van Vlierberghe, P.; De Moerloose, B. Molecular basis and clinical significance of genetic aberrations in B-cell precursor acute lymphoblastic leukemia. Exp. Hematol. 2015, 43, 640–653. [Google Scholar] [CrossRef]
  98. Kubota-Tanaka, M.; Osumi, T.; Miura, S.; Tsujimoto, H.; Imamura, T.; Nishimura, A.; Oki, K.; Nakamura, K.; Miyamoto, S.; Inoue, K.; et al. B-lymphoblastic lymphoma with TCF3-PBX1 fusion gene. Haematologica 2019, 104, e35–e37. [Google Scholar] [CrossRef]
  99. Rowsey, R.A.; Smoley, S.A.; Williamson, C.M.; Vasmatzis, G.; Smadbeck, J.B.; Ning, Y.; Greipp, P.T.; Hoppman, N.L.; Baughn, L.B.; Ketterling, R.P.; et al. Characterization of TCF3 rearrangements in pediatric B-lymphoblastic leukemia/lymphoma by mate-pair sequencing (MPseq) identifies complex genomic rearrangements and a novel TCF3/TEF gene fusion. Blood Cancer J. 2019, 9, 81. [Google Scholar] [CrossRef]
  100. Tirado, C.A.; Shabsovich, D.; Yeh, L.; Pullarkat, S.T.; Yang, L.; Kallen, M.; Rao, N. A (1;19) translocation involving TCF3-PBX1 fusion within the context of a hyperdiploid karyotype in adult B-ALL: A case report and review of the literature. Biomark. Res. 2015, 3, 4. [Google Scholar] [CrossRef]
  101. Lehnertz, B.; Chagraoui, J.; MacRae, T.; Tomellini, E.; Corneau, S.; Mayotte, N.; Boivin, I.; Durand, A.; Gracias, D.; Sauvageau, G. HLF expression defines the human hematopoietic stem cell state. Blood 2021, 138, 2642–2654. [Google Scholar] [CrossRef]
  102. Paulsson, K.; Jonson, T.; Øra, I.; Olofsson, T.; Panagopoulos, I.; Johansson, B. Characterisation of genomic translocation breakpoints and identification of an alternative TCF3/PBX1 fusion transcript in t(1;19)(q23;p13)-positive acute lymphoblastic leukaemias. Br. J. Haematol. 2007, 138, 196–201. [Google Scholar] [CrossRef]
  103. Lin, N.; Yan, X.; Cai, D.; Wang, L. Leukemia With TCF3-ZNF384 Rearrangement as a Distinct Subtype of Disease With Distinct Treatments: Perspectives From A Case Report and Literature Review. Front. Oncol. 2021, 11, 709036. [Google Scholar] [CrossRef] [PubMed]
  104. Fischer, U.; Forster, M.; Rinaldi, A.; Risch, T.; Sungalee, S.; Warnatz, H.-J.; Bornhauser, B.; Gombert, M.; Kratsch, C.; Stütz, A.M.; et al. Genomics and drug profiling of fatal TCF3-HLFâ ’positive acute lymphoblastic leukemia identifies recurrent mutation patterns and therapeutic options. Nat. Genet. 2015, 47, 1020–1029. [Google Scholar] [CrossRef]
  105. Takeda, R.; Yokoyama, K.; Ogawa, M.; Kawamata, T.; Fukuyama, T.; Kondoh, K.; Takei, T.; Nakamura, S.; Ito, M.; Yusa, N.; et al. The first case of elderly TCF3-HLF-positive B-cell acute lymphoblastic leukemia. Leuk Lymphoma 2019, 60, 2821–2824. [Google Scholar] [CrossRef]
  106. Lejman, M.; Włodarczyk, M.; Zawitkowska, J.; Kowalczyk, J.R. Comprehensive chromosomal aberrations in a case of a patient with TCF3-HLF-positive BCP-ALL. BMC Med. Genom. 2020, 13, 58. [Google Scholar] [CrossRef]
  107. Mouttet, B.; Vinti, L.; Ancliff, P.; Bodmer, N.; Brethon, B.; Cario, G.; Chen-Santel, C.; Elitzur, S.; Hazar, V.; Kunz, J.; et al. Durable remissions in TCF3-HLF positive acute lymphoblastic leukemia with blinatumomab and stem cell transplantation. Haematologica 2019, 104, e244–e247. [Google Scholar] [CrossRef] [PubMed]
  108. Wang, T.; Wan, X.; Yang, F.; Shi, W.; Liu, R.; Ding, L.; Tang, Y.; Luo, C.; Yang, X.; Ma, Y.; et al. Successful Treatment of TCF3-HLF–positive Childhood B-ALL with Chimeric Antigen Receptor T-Cell Therapy. Clin. Lymphoma Myeloma Leuk. 2021, 21, 386–392. [Google Scholar] [CrossRef] [PubMed]
  109. Li, Z.; Lee, S.H.R.; Ni Chin, W.H.; Lu, Y.; Jiang, N.; Lim, E.H.H.; Coustan-Smith, E.; Chiew, K.H.; Oh, B.L.Z.; Koh, G.S.; et al. Distinct clinical characteristics of DUX4- And PAX5-altered childhood B-lymphoblastic leukemia. Blood Adv. 2021, 5, 5226–5238. [Google Scholar] [CrossRef]
  110. Paietta, E.; Roberts, K.G.; Wang, V.; Gu, Z.; Buck, G.A.N.; Pei, D.; Cheng, C.; Levine, R.L.; Abdel-Wahab, O.; Cheng, Z.; et al. Molecular classification improves risk assessment in adult BCR-ABL1–negative B-ALL. Blood 2021, 138, 948. [Google Scholar] [CrossRef]
  111. Yasuda, T.; Tsuzuki, S.; Kawazu, M.; Hayakawa, F.; Kojima, S.; Ueno, T.; Imoto, N.; Kohsaka, S.; Kunita, A.; Doi, K.; et al. Recurrent DUX4 fusions in B cell acute lymphoblastic leukemia of adolescents and young adults. Nat. Genet. 2016, 48, 569–574. [Google Scholar] [CrossRef]
  112. Liu, Y.-F.; Wang, B.-Y.; Zhang, W.-N.; Huang, J.-Y.; Li, B.-S.; Zhang, M.; Jiang, L.; Li, J.-F.; Wang, M.-J.; Dai, Y.-J.; et al. Genomic Profiling of Adult and Pediatric B-cell Acute Lymphoblastic Leukemia. EBioMedicine 2016, 8, 173–183. [Google Scholar] [CrossRef]
  113. Zhang, J.; McCastlain, K.; Yoshihara, H.; Xu, B.; Chang, Y.; Churchman, M.L.; Wu, G.; Li, Y.; Wei, L.; Iacobucci, I.; et al. Deregulation of DUX4 and ERG in acute lymphoblastic leukemia. Nat. Genet. 2016, 48, 1481–1489. [Google Scholar] [CrossRef]
  114. Lilljebjörn, H.; Fioretos, T. New oncogenic subtypes in pediatric B-cell precursor acute lymphoblastic leukemia. Blood 2017, 130, 1395–1401. [Google Scholar] [CrossRef] [PubMed]
  115. Hirabayashi, S.; Butler, E.; Ohki, K.; Kiyokawa, N.; Bergmann, A.K.; Boer, J.M.; Cavé, H.; Cazzaniga, G.; Yeoh, A.E.J.; Imamura, T.; et al. Acute Lymphoblastic Leukemia with Zinc-Finger Protein 384 (ZNF384)-Related Rearrangements: A Retrospective Analysis from the Ponte Di Legno Childhood ALL Working Group. Blood 2019, 134, 652. [Google Scholar] [CrossRef]
  116. Qin, Y.-Z.; Jiang, Q.; Xu, L.-P.; Wang, Y.; Jiang, H.; Dao, F.-T.; Chen, W.-M.; Zhao, X.-S.; Liu, Y.-R.; Zhang, X.-H.; et al. The Prognostic Significance of ZNF384 Fusions in Adult Ph-Negative B-Cell Precursor Acute Lymphoblastic Leukemia: A Comprehensive Cohort Study From a Single Chinese Center. Front. Oncol. 2021, 11, 632532. [Google Scholar] [CrossRef] [PubMed]
  117. Zhu, L.; Bai, W.; Cheng, Q.; Fang, J. ZNF384-Related Fusion Genes in Acute Lymphoblastic Leukemia. Cancer Control 2023, 30, 1–7. [Google Scholar] [CrossRef]
  118. Roberts, K.G.; Mullighan, C.G. The biology of B-progenitor acute lymphoblastic leukemia. Cold Spring Harb. Perspect. Med. 2020, 10, 1–22. [Google Scholar] [CrossRef] [PubMed]
  119. Shinsuke, H.; Kentaro, O.; Kazuhiko, N.; Hitoshi, I.; Yukihide, M.; Kohji, O.; Akinori, Y.; Kazuki, T.; Yuya, S.; Ai, Y.; et al. ZNF384-related fusion genes define a subgroup of childhood B-cell precursor acute lymphoblastic leukemia with a characteristic immunotype. Haematologica 2017, 102, 118–129. [Google Scholar] [CrossRef]
  120. Heavey, B.; Charalambous, C.; Cobaleda, C.; Busslinger, M. Myeloid lineage switch of Pax5 mutant but not wild-type B cell progenitors by C/EBPα and GATA factors. EMBO J. 2003, 22, 3887–3897. [Google Scholar] [CrossRef]
  121. Kansal, R. Diagnosis and Molecular Pathology of Lymphoblastic Leukemias and Lymphomas in the Era of Genomics and Precision Medicine: Historical Evolution and Current Concepts—Part 2: B-/T-Cell Acute Lymphoblastic Leukemias. Lymphatics 2023, 1, 118–154. [Google Scholar] [CrossRef]
  122. Qian, M.; Zhang, H.; Kham, S.K.-Y.; Liu, S.; Jiang, C.; Zhao, X.; Lu, Y.; Goodings, C.; Lin, T.-N.; Zhang, R.; et al. Whole-transcriptome sequencing identifies a distinct subtype of acute lymphoblastic leukemia with predominant genomic abnormalities of EP300 and CREBBP. Genome Res. 2017, 27, 185–195. [Google Scholar] [CrossRef]
  123. Ohki, K.; Kiyokawa, N.; Saito, Y.; Hirabayashi, S.; Nakabayashi, K.; Ichikawa, H.; Momozawa, Y.; Okamura, K.; Yoshimi, A.; Ogata-Kawata, H.; et al. Clinical and molecular characteristics of MEF2D fusion-positive B-cell precursor acute lymphoblastic leukemia in childhood, including a novel translocation resulting in MEF2D-HNRNPH1 gene fusion. Haematologica 2019, 104, 128–137. [Google Scholar] [CrossRef] [PubMed]
  124. Gu, Z.; Churchman, M.; Roberts, K.; Li, Y.; Liu, Y.; Harvey, R.C.; McCastlain, K.; Reshmi, S.C.; Payne-Turner, D.; Iacobucci, I.; et al. Genomic analyses identify recurrent MEF2D fusions in acute lymphoblastic leukaemia. Nat. Commun. 2016, 7, 13331. [Google Scholar] [CrossRef] [PubMed]
  125. Ohki, K.; Butler, E.R.; Kiyokawa, N.; Hirabayashi, S.; Bergmann, A.K.; Möricke, A.; Boer, J.M.; Cavé, H.; Cazzaniga, G.; Yeoh, A.E.J.; et al. Clinical characteristics and outcomes of B-cell precursor ALL with MEF2D rearrangements: A retrospective study by the Ponte di Legno Childhood ALL Working Group. Leukemia 2023, 37, 212–216. [Google Scholar] [CrossRef]
  126. Leongamornlert, D.; Gutiérrez-Abril, J.; Lee, S.; Barretta, E.; Creasey, T.; Gundem, G.; Levine, M.F.; Arango-Ossa, J.E.; Liosis, K.; Medina-Martinez, J.S.; et al. Diagnostic utility of whole genome sequencing in adults with B-other acute lymphoblastic leukemia. Blood Adv. 2023, 7, 3862–3873. [Google Scholar] [CrossRef]
  127. Van Outersterp, I.; Hormann, F.M.; Hoogkamer, A.Q.; Boeree, A.; Broek, S.A.V.D.; Boer, M.L.D.; Boer, J.M. Characterization of a novel MEF2D-BCL9 fusion-positive acute lymphoblastic leukemia cell line. Haematologica 2023, 108, 2859–2864. [Google Scholar] [CrossRef]
  128. Zhang, M.; Zhang, H.; Li, Z.; Bai, L.; Wang, Q.; Li, J.; Jiang, M.; Xue, Q.; Cheng, N.; Zhang, W.; et al. Functional, structural, and molecular characterizations of the leukemogenic driver MEF2D-HNRNPUL1 fusion. Blood 2022, 140, 1390–1407. [Google Scholar] [CrossRef] [PubMed]
  129. Boer, J.M.; Valsecchi, M.G.; Hormann, F.M.; Antić, Ž.; Zaliova, M.; Schwab, C.; Cazzaniga, G.; Arfeuille, C.; Cavé, H.; Attarbaschi, A.; et al. Favorable outcome of NUTM1-rearranged infant and pediatric B cell precursor acute lymphoblastic leukemia in a collaborative international study. Leukemia 2021, 35, 2978–2982. [Google Scholar] [CrossRef]
  130. Hormann, F.M.; Hoogkamer, A.Q.; Beverloo, H.B.; Boeree, A.; Dingjan, I.; Wattel, M.M.; Stam, R.W.; Escherich, G.; Pieters, R.; den Boer, M.L.; et al. NUTM1 is a recurrent fusion gene partner in B-cell precursor acute lymphoblastic leukemia associated with increased expression of genes on chromosome band 10p12.31-12.2. Haematologica 2019, 104, 455–5459. [Google Scholar] [CrossRef]
  131. Charlab, R.; Racz, R. The expanding universe of NUTM1 fusions in pediatric cancer. Clin. Transl. Sci. 2023, 16, 1331–1339. [Google Scholar] [CrossRef]
  132. Boer, J.; Valsecchi, M.; Hormann, F.; Antic, Z.; Zaliova, M.; Schwab, C.; Cazzaniga, G.; Arfeuille, C.; Cavé, H.; Attarbaschi, A.; et al. NUTM1-Rearranged Infant and Pediatric B Cell Precursor Acute Lymphoblastic Leukemia: A Good Prognostic Subtype Identified in a Collaborative International Study. Blood 2020, 136, 25–26. [Google Scholar] [CrossRef]
  133. McClure, B.J.; Pal, M.; Heatley, S.L.; Rehn, J.; Schutz, C.; Breen, J.; Venn, N.C.; Sutton, R.; Khaw, S.L.; Yeung, D.T.; et al. Two novel cases of NUTM1-rearranged B-cell acute lymphoblastic leukaemia presenting with high-risk features. Br. J. Haematol. 2022, 196, 1407–1411. [Google Scholar] [CrossRef] [PubMed]
  134. Kimura, S.; Montefiori, L.; Iacobucci, I.; Zhao, Y.; Gao, Q.; Paietta, E.M.; Haferlach, C.; Laird, A.D.; Mead, P.E.; Gu, Z.; et al. Enhancer retargeting of CDX2 and UBTF::ATXN7L3 define a subtype of high-risk B-progenitor acute lymphoblastic leukemia. Blood 2022, 136, 3519–3531. [Google Scholar] [CrossRef] [PubMed]
  135. Passet, M.; Kim, R.; Gachet, S.; Sigaux, F.; Chaumeil, J.; Galland, A.; Sexton, T.; Quentin, S.; Hernandez, L.; Larcher, L.; et al. Concurrent CDX2 cis-deregulation and UBTF::ATXN7L3 fusion define a novel high-risk subtype of B-cell ALL. Blood 2022, 136, 3505–3518. [Google Scholar] [CrossRef] [PubMed]
  136. Yang, M.; Safavi, S.; Woodward, E.L.; Duployez, N.; Olsson-Arvidsson, L.; Ungerbäck, J.; Sigvardsson, M.; Zaliova, M.; Zuna, J.; Fioretos, T.; et al. 13q12.2 deletions in acute lymphoblastic leukemia lead to upregulation of FLT3 through enhancer hijacking. Blood 2020, 136, 946–956. [Google Scholar] [CrossRef] [PubMed]
  137. Bastian, L.; Hartmann, A.M.; Beder, T.; Hänzelmann, S.; Kässens, J.; Bultmann, M.; Hoeppner, M.P.; Franzenburg, S.; Wittig, M.; Franke, A.; et al. UBTF::ATXN7L3 gene fusion defines novel B cell precursor ALL subtype with CDX2 expression and need for intensified treatment. Leukemia 2022, 36, 1676–1680. [Google Scholar] [CrossRef]
  138. Galland, A.; Gourain, V.; Habbas, K.; Güler, Y.; Martin, E.; Ebel, C.; Tavian, M.; Vallat, L.; Chenard, M.; Mauvieux, L.; et al. CDX2 expression in the hematopoietic lineage promotes leukemogenesis via TGFβ inhibition. Mol. Oncol. 2021, 15, 2318–2329. [Google Scholar] [CrossRef]
  139. Darvishi, M.; Mashati, P.; Khosravi, A. The clinical significance of CDX2 in leukemia: A new perspective for leukemia research. Leuk. Res. 2018, 72, 45–51. [Google Scholar] [CrossRef]
  140. Jung, M.; Schieck, M.; Hofmann, W.; Tauscher, M.; Lentes, J.; Bergmann, A.; Stelter, M.; Möricke, A.; Alten, J.; Schlegelberger, B.; et al. Frequency and prognostic impact of PAX5 p.P80R in pediatric acute lymphoblastic leukemia patients treated on an AIEOP-BFM acute lymphoblastic leukemia protocol. Genes. Chromosomes Cancer 2020, 59, 667–671. [Google Scholar] [CrossRef]
  141. Jia, Z.; Gu, Z. PAX5 alterations in B-cell acute lymphoblastic leukemia. Front. Oncol. 2022, 12, 1023606. [Google Scholar] [CrossRef]
  142. Gu, Z.; Churchman, M.L.; Roberts, K.G.; Moore, I.; Zhou, X.; Nakitandwe, J.; Hagiwara, K.; Pelletier, S.; Gingras, S.; Berns, H.; et al. PAX5-driven subtypes of B-progenitor acute lymphoblastic leukemia. Nat. Genet. 2019, 51, 296–307. [Google Scholar] [CrossRef]
  143. Bastian, L.; Schroeder, M.P.; Eckert, C.; Schlee, C.; Tanchez, J.O.; Kämpf, S.; Wagner, D.L.; Schulze, V.; Isaakidis, K.; Lázaro-Navarro, J.; et al. PAX5 biallelic genomic alterations define a novel subgroup of B-cell precursor acute lymphoblastic leukemia. Leukemia 2019, 33, 1895–1909. [Google Scholar] [CrossRef]
  144. Østergaard, A.; Boer, J.M.; van Leeuwen, F.N.; Pieters, R.; Den Boer, M.L. IKZF1 in acute lymphoblastic leukemia: The rise before the fall? Leuk. Lymphoma 2024, 29, 2396046. [Google Scholar] [CrossRef] [PubMed]
  145. Stanulla, M.; Cavé, E.E.; Cavé, C.; Moorman, A.V. IKZF1 deletions in pediatric acute lymphoblastic leukemia: Still a poor prognostic marker? Blood 2020, 135, 252–260. [Google Scholar] [CrossRef] [PubMed]
  146. Tsai, H.K.; Gogakos, T.; Lip, V.; Tsai, J.; Li, Y.D.; Fisch, A.; Weiss, J.; Grimmett, L.; Tran, T.H.; Caron, M.; et al. Outlier expression of isoforms by targeted RNA sequencing as clinical markers of genomic variants in B lymphoblastic leukemia and other tumor types. medRxiv 2022. [Google Scholar] [CrossRef]
  147. Fazio, G.; Biondi, A.; Cazzaniga, G. The Role of PAX5 in ALL. Nov. Asp. Acute Lymphoblstic Leuk. 2011, 1, 211–234. [Google Scholar]
  148. Fazio, G.; Bresolin, S.; Silvestri, D.; Quadri, M.; Saitta, C.; Vendramini, E.; Buldini, B.; Palmi, C.; Bardini, M.; Grioni, A.; et al. PAX5 fusion genes are frequent in poor-risk childhood acute lymphoblastic leukaemia and can be targeted with BIBF1120. EBioMedicine 2022, 83, 104224. [Google Scholar] [CrossRef]
  149. Schwab, C.; Nebral, K.; Chilton, L.; Leschi, C.; Waanders, E.; Boer, J.M.; Žaliová, M.; Sutton, R.; Öfverholm, I.I.; Ohki, K.; et al. Intragenic amplification of PAX5: A novel subgroup in B-cell precursor acute lymphoblastic leukemia? Blood Adv. 2017, 1, 1473–1477. [Google Scholar] [CrossRef]
  150. Jean, J.; Kovach, A.E.; Doan, A.; Oberley, M.J.; Ji, J.; Schmidt, R.J.; Biegel, J.A.; Bhojwani, D.; Raca, G. Characterization of PAX5 intragenic tandem multiplication in pediatric B-lymphoblastic leukemia by optical genome mapping. Blood Adv. 2022, 6, 3343–3346. [Google Scholar] [CrossRef]
  151. He, L.; Li, Z.; Fan, X.; Chen, J.; Wu, H.; Fu, Y. ‘Double hit’ B-lymphoblastic lymphoma with concurrent IGH/BCL2 and 8q24/MYC translocations: A case report. Transl. Cancer Res. 2021, 10, 1594–1598. [Google Scholar] [CrossRef]
  152. Genescà, E.; la Starza, R. Early T-Cell Precursor ALL and Beyond: Immature and Ambiguous Lineage T-ALL Subsets. Cancers 2022, 14, 1873. [Google Scholar] [CrossRef]
  153. Coustan-Smith, E.; Mullighan, C.G.; Onciu, M.; Behm, F.G.; Raimondi, S.C.; Pei, D.; Cheng, C.; Su, X.; Rubnitz, J.E.; Basso, G.; et al. Early T-cell precursor leukaemia: A subtype of very high-risk acute lymphoblastic leukaemia. Lancet Oncol. 2009, 10, 147–156. [Google Scholar] [CrossRef] [PubMed]
  154. Zhang, J.; Ding, L.; Holmfeldt, L.; Wu, G.; Heatley, S.L.; Payne-Turner, D.; Easton, J.; Chen, X.; Wang, J.; Rusch, M.; et al. The genetic basis of early T-cell precursor acute lymphoblastic leukaemia. Nature 2012, 481, 157–163. [Google Scholar] [CrossRef] [PubMed]
  155. Jain, N.; Lamb, A.V.; O’brien, S.; Ravandi, F.; Konopleva, M.; Jabbour, E.; Zuo, Z.; Jorgensen, J.; Lin, P.; Pierce, S.; et al. Early T-cell precursor acute lymphoblastic leukemia/lymphoma (ETP-ALL/LBL) in adolescents and adults: A high-risk subtype. Blood 2016, 127, 1863–1869. [Google Scholar] [CrossRef] [PubMed]
  156. Patel, S.; Nayernama, A.; Jones, S.C.; de Claro, R.A.; Waldron, P.E. A novel germline variant in PIK3R1 results in SHORT syndrome associated withTAL/LMO T-cell acute lymphoblastic leukemia. Am. J. Hematol. 2020, 95, E335–E336. [Google Scholar] [CrossRef]
  157. Bardelli, V.; Arniani, S.; Pierini, V.; Di Giacomo, D.; Pierini, T.; Gorello, P.; Mecucci, C.; La Starza, R. T-cell acute lymphoblastic leukemia: Biomarkers and their clinical usefulness. Genes 2021, 12, 1118. [Google Scholar] [CrossRef]
  158. Tasian, S.K.; Hunger, S.P. Genomic characterization of paediatric acute lymphoblastic leukaemia: An opportunity for precision medicine therapeutics. Br. J. Haematol. 2017, 176, 867–882. [Google Scholar] [CrossRef]
  159. Mroczek, A.; Zawitkowska, J.; Kowalczyk, J.; Lejman, M. Comprehensive overview of gene rearrangements in childhood T-cell acute lymphoblastic leukaemia. Int. J. Mol. Sci. 2021, 22, 808. [Google Scholar] [CrossRef]
  160. Papenhausen, P.; Kelly, C.A.; Zhang, Z.; Tepperberg, J.; Burnside, R.D.; Schwartz, S. Multidisciplinary analysis of pediatric T-ALL: 9q34 gene fusions. Cancer Genet. 2019, 1, 231–232. [Google Scholar] [CrossRef]
  161. Liu, Y.; Easton, J.; Shao, Y.; Maciaszek, J.; Wang, Z.; Wilkinson, M.R.; McCastlain, K.; Edmonson, M.; Pounds, S.B.; Shi, L.; et al. The genomic landscape of pediatric and young adult T-lineage acute lymphoblastic leukemia. Nat. Genet. 2017, 49, 1211–1218. [Google Scholar] [CrossRef]
  162. Mansur, M.B.; Furness, C.L.; Nakjang, S.; Enshaei, A.; Alpar, D.; Colman, S.M.; Minto, L.; Irving, J.; Poole, B.V.; Noronha, E.P.; et al. The genomic landscape of teenage and young adult T-cell acute lymphoblastic leukemia. Cancer Med. 2021, 10, 4864–4873. [Google Scholar] [CrossRef]
  163. Wang, H.; Zhou, Y.; Huang, X.; Zhang, Y.; Qian, J.; Li, J.; Li, X.; Li, C.; Lou, Y.; Mai, W.; et al. CDKN2A deletions are associated with poor outcomes in 101 adults with T-cell acute lymphoblastic leukemia. Am. J. Hematol. 2021, 96, 312–319. [Google Scholar] [CrossRef] [PubMed]
  164. De Bie, J.; Quessada, J.; Tueur, G.; Lefebvre, C.; Luquet, I.; Toujani, S.; Cuccuini, W.; Lafage-Pochitaloff, M.; Michaux, L. Cytogenetics in the management of T-cell acute lymphoblastic leukemia (T-ALL): Guidelines from the Groupe Francophone de Cytogénétique Hématologique (GFCH). Curr. Res. Transl. Med. 2023, 71, 103431. [Google Scholar] [CrossRef]
  165. Coustan-Smith, E.; Conter, V. Closing the circle for ETP ALL. Blood 2023, 142, 2039–2040. [Google Scholar] [CrossRef] [PubMed]
  166. Sin, C.F.; Man, P.H.M. Early T-Cell Precursor Acute Lymphoblastic Leukemia: Diagnosis, Updates in Molecular Pathogenesis, Management, and Novel Therapies. Front. Oncol. 2021, 11, 750789. [Google Scholar] [CrossRef]
  167. Li, Z.; Song, Y.; Zhang, Y.; Li, C.; Wang, Y.; Xue, W.; Lu, L.; Jin, M.; Zhou, Z.; Wang, X.; et al. Genomic and outcome analysis of adult T-cell lymphoblastic lymphoma. Haematologica 2020, 105, E107–E110. [Google Scholar] [CrossRef]
  168. Khabirova, E.; Jardine, L.; Coorens, T.H.H.; Webb, S.; Treger, T.D.; Engelbert, J.; Porter, T.; Prigmore, E.; Collord, G.; Piapi, A.; et al. Single-cell transcriptomics reveals a distinct developmental state of KMT2A-rearranged infant B-cell acute lymphoblastic leukemia. Nat. Med. 2022, 28, 743–751. [Google Scholar] [CrossRef]
  169. Iacobucci, I.; Witkowski, M.T.; Mullighan, C.G. Single-cell analysis of acute lymphoblastic and linage-ambiguous leukemia: Approaches and molecular insights. Blood 2023, 141, 356–368. [Google Scholar] [CrossRef]
  170. Zhang, J.; Zeng, L.; Wang, Y.; Pan, J.; Li, X.; Feng, B.; Yang, Q. Gene Mutations Related to Glucocorticoid Resistance in Pediatric Acute Lymphoblastic Leukemia. Front. Pediatr. 2022, 10, 831229. [Google Scholar] [CrossRef] [PubMed]
  171. Waanders, E.; Gu, Z.; Dobson, S.M.; Antić, Ž.; Crawford, J.C.; Ma, X.; Edmonson, M.N.; Payne-Turner, D.; van de Vorst, M.; Jongmans, M.C.; et al. Mutational Landscape and Patterns of Clonal Evolution in Relapsed Pediatric Acute Lymphoblastic Leukemia. Blood Cancer Discov. 2020, 1, 96–111. [Google Scholar] [CrossRef]
  172. van der Ham, C.G.; Suurenbroek, L.C.; Kleisman, M.M.; Antić, Ž.; Lelieveld, S.H.; Yeong, M.; Westera, L.; Sonneveld, E.; Hoogerbrugge, P.M.; van der Velden, V.H.J.; et al. Mutational mechanisms in multiply relapsed pediatric acute lymphoblastic leukemia. Leukemia 2024, 38, 2366–2375. [Google Scholar] [CrossRef]
  173. Ching, T.; Duncan, M.E.; Newman-Eerkes, T.; McWhorter, M.M.E.; Tracy, J.M.; Steen, M.S.; Brown, R.P.; Venkatasubbarao, S.; Akers, N.K.; Vignali, M.; et al. Analytical evaluation of the clonoSEQ Assay for establishing measurable (minimal) residual disease in acute lymphoblastic leukemia, chronic lymphocytic leukemia, and multiple myeloma. BMC Cancer 2020, 20, 612. [Google Scholar] [CrossRef] [PubMed]
  174. Gil, J.V.; Such, E.; Sargas, C.; Simarro, J.; Miralles, A.; Pérez, G.; de Juan, I.; Palanca, S.; Avetisyan, G.; Santiago, M.; et al. Design and Validation of a Custom Next-Generation Sequencing Panel in Pediatric Acute Lymphoblastic Leukemia. Int. J. Mol. Sci. 2023, 24, 4440. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Milestones in ALL based on the use and implementation of molecular cytogenetic and NGS technologies. GEP: gene expression profile, LOH: loss of heterozygosity, CNAs: copy number alterations, iAMP: intrachromosomal amplification, SNP: single nucleotide polymorphism, GWAS: genome-wide association studies, WGS: whole-genome sequencing, WHO: World Health Organization, MTX: methotrexate, MRD: minimal residual disease, tNGS: targeted NGS.
Figure 1. Milestones in ALL based on the use and implementation of molecular cytogenetic and NGS technologies. GEP: gene expression profile, LOH: loss of heterozygosity, CNAs: copy number alterations, iAMP: intrachromosomal amplification, SNP: single nucleotide polymorphism, GWAS: genome-wide association studies, WGS: whole-genome sequencing, WHO: World Health Organization, MTX: methotrexate, MRD: minimal residual disease, tNGS: targeted NGS.
Cancers 16 03965 g001
Figure 2. Recent ALL classification and risk subtypes based on genomic/transcriptomic profiles in child patients. As can be seen, genomic classification not only contributes to a better diagnosis but also has prognostic implications. Considerations: The width of the band corresponds to the percentage of cases with each of the alterations. An * indicates the provisional entity status according to the International Consensus Classification (2022).
Figure 2. Recent ALL classification and risk subtypes based on genomic/transcriptomic profiles in child patients. As can be seen, genomic classification not only contributes to a better diagnosis but also has prognostic implications. Considerations: The width of the band corresponds to the percentage of cases with each of the alterations. An * indicates the provisional entity status according to the International Consensus Classification (2022).
Cancers 16 03965 g002
Figure 3. Frequency of recurrent alterations in T-ALL. Based on average values found in the reports analyzed in this review.
Figure 3. Frequency of recurrent alterations in T-ALL. Based on average values found in the reports analyzed in this review.
Cancers 16 03965 g003
Table 1. Recent findings in ALL based on NGS applications.
Table 1. Recent findings in ALL based on NGS applications.
YearNGSSample
Size
ApplicationMain Findings
2017WES1000 cases (B-ALL and T-ALL)Therapeutic targets IdentificationClinical and molecular data at diagnosis and remission of ALL patients, including MRD are currently being collected. They will include the identification of somatic or germline variants associated with response to treatment, assessment of diversity, clonal evolution, etc. This study proposes to use NGS in diagnosis alongside morphological, immunophenotyping, and other molecular approaches, including scNGS [35].
2018RNAseq1223 B-ALL samplesCharacterization of molecular subtypes for diagnostic accuracy and risk estimationThis study proposes fourteen molecular subtypes based on transcriptional profiling, including six previously undescribed subtypes: (i) PAX5 or CRLF2 fusions, (ii) PAX5 mutation (p.P80R), (iii) IKZF1 mutation (p.N159Y), (iv) ZEB2 mutation (p.H1038R) or IGH/CEBPE fusion, (v) HLF rearrangements, and (vi) NUTM1 rearrangements [36].
2019RNAseq170 B-ALL samplesBiomarker identificationReport on the first oncogenic subtype defined by a point mutation in ALL (PAX5 P80R) that may harbor a second PAX5 allele inactivation, biallelic CDKN2A deletion, and RAS signaling alterations [37].
2019tNGS17 childhood B-ALL samplesTranslational research for diagnosis and monitoring of ALL patientsMatched diagnostic, remission, and relapse samples allowed for the detection of “potentially actionable gene variants” such as CDKN2A/B deletions and KRAS, ABL1, FLT3, and NOTCH mutations, as well as the detection of germline alterations such as TP53 R248Q [38].
2019RNA-seq40 diagnoses and 42 relapsed BCP-ALL samplesInsights into epigenomic changesAt diagnosis, DUX4-specific lncRNAs were associated with the TGF-beta and p53 signaling pathways. For the Ph-like subtype, lncRNAs were correlated with genes involved in the JAK-STAT and PIK3/AKT/mTOR signaling pathways. For all subtypes, lncRNAs were found to be differentially expressed at diagnosis and relapse. About 80% of the lncRNAs were downregulated at relapse, e.g., TCL6 in the DUX4r subtype, and, LINC00312 and miR-17-92a-1 in the DUX4r and Ph-like subtypes [39].
2020WGS,
ChIP-seq
RNA-seq
RRBS-Seq
34 B-ALL
Samples
Epigenome in clonal evolution and
treatment failure
This study provides the first reports of the role of epigenetic mechanisms in driving clonal evolution in B-ALL. Changes in gene expression were directly correlated with epigenetic alterations with changes in H3K27ac, and PRC2 (Polycomb complex) hypermethylation was detected at relapse in many patients [40].
2020ATAC-Seq
RNA-Seq
Primary cells of T-ALL patients and patient-derived xenograft modelsInsights into epigenomic changesTwo distinct groups of T-ALL were identified based on accessible transcription factor (TF)-binding-motifs profiles. The first group was characterized by highly accessible SPI1-binding sites, and the second lacked hyper-accessible SPI1-binding sites and DAB1 expression [41].
2021RNAseq
tNGS
58 AYAs and adult B-ALL samplesALL diagnosis and stratificationThis study introduces the first custom panel (LYmphoid Next-Generation Sequencing-LYNX Panel) as a versatile tool for the simultaneous analysis of genomic markers (mutations, indels, chromosomal aberrations, and IG/TR rearrangements) in most common lymphoid neoplasms (70 lymphoma-related genes) with a sensitivity of 5% variant allele frequency, as well as reliable identification of large genome-wide (≥6 Mb) and recurrent chromosomal aberrations (≥300 kb). This panel allowed for the description of novel variants classified as pathogenic or probable pathogenic in TP53, NRAS, JAK2, PAX5, CREBBP, NF1, FLT3, ATM, KRAS, and RUNX1 genes, as well as IKZF1plus profile identification [42,43].
2021in silico analysis1978 samples of childhood B-ALLBiomarker identification and prognostic associationADAM6 may be a novel genetic biomarker. Homozygous ADAM6 deletion was detected around 60% of cases and was associated with higher total leukocyte count, higher first relapse rate, MRD-positive status, and higher mutation variants. It was more common in the TCF3::PBX1 and ETV6::RUNX1 subtypes [44].
2021DNA-seq
ddPCR
376 LPDs and
14 LPD cell lines
Technique Performance/Translational researchAssays were designed to capture the diversity and functional variation of joining genes across 1243 regions included in a tNGS panel. Sensitivity for detecting IG/TCR clonality ranged from 96.5 to 97.3%. The detection of CNAs was dependent on the locus, with high sensitivities being reported for trisomy 12 and del(11q) and lower sensitivities for del(17p) and del(13q) [45].
2021DNA-seq
RNA-seq
165 ALL patients at diagnosisTechnique
performance/translational research
EuroClonality-NGS captures validation as a tNGS assay for SNVs, CNVs, Indels, Translocations, and IGH and TRC rearrangements. Less than half of the markers were also found in the RNAseq data (44%). The assay performance for somatic mutations was 100% reproducible for both sensitivity and specificity at >4% VAF [46].
2022WGBS
WGS
WES
82
B-ALL patients
Insights into the epigenome Local hypermethylation was observed in the global DNA methylation landscape of ALL and, more specifically, in T-ALL patients [47].
2022Methyla-tionEPIC BeadChip Arrays38 children with B-ALL pediatrics cases and controlsMethylation profile and microRNA changes landscape Among specific genetic subtypes (ETV6::RUNX1, TCF3::PBX1, IGH rearrangement, hyperdiploidy), miRNAs with differentially methylated sites were found. The results showed that MIR326, MIR200c, MIR125B, MIR203, and MIR181A were significantly differentially expressed in B-ALL cases compared with healthy controls. This study was complemented by meta-analysis and in silico analysis to identify differentially expressed miRNAs in B-ALL. Although no consensus dataset was found among the studies, miR-181b, miR-181a, miR-128, miR-128a, miR-181c, miR-155, miR-142-3p, and miR-451 were postulated as biomarkers associated with B-ALL [48,49].
2023ATAC-Seq
ChIP-Seq
Hi-C
154 B-ALL Insights into epigenomic changesThis was the first large-scale analysis of chromatin accessibility in the B-ALL genome across extensive B-ALL subtypes. The percentage of differentially accessible chromatin sites (DAS) was associated with each molecular subtype. It was higher in the KMT2Ar, TCF3-PBX1 fusion, and ZNF384r subtypes. Analysis of bound TF motif footprint prevalence identified several ETS family TFs (EHF, ELF3, SPI1, and SPIB), zinc finger TFs (ZNF263, ZNF460, ZNF740, and ZNF148), and CTCF as the most altered motifs leading to differences in chromatin accessibility [50].
Subtypes: BCR::ABL1, DUX4r, ETV6::RUNX1, high hyperdiploidy, low hypodiploidy, KMT2Ar, Ph-like, PAX5alt, TCF3::PBX1, ZNF384r, B-other samples, and B-ALL cell lines (697, BALL1, Nalm6, REH, RS411, SEM, and SUPB15)
TGS: targeted next-generation sequencing. scNGS: single-cell next-generation sequencing. WES: whole-exome sequencing. WGBS: whole-genome bisulfite sequencing. DAS: differentially accessible chromatin sites. MRD: minimal residual disease. scRRBS/RRBS-Seq: reduced-representation bisulfite sequencing or single-cell reduced-representation bisulfite sequencing. Hi-C: 3C-Seq to map chromatin contacts genome-wide or chromosome conformation capture. ddPCR: droplet digital PCR. DNAamp IG/TR assays developed by EuroClonality-NGS (DNA-amplicon-based methods). HeH: high hyperdiploidy. Potentially actionable genes include gene alterations that could lead to a newer molecular-targeted therapy, a change in the selection of drugs, or a change in patient or family counseling and management by identifying germline cancer-predisposing gene alterations. LPDs: lymphoproliferative disorders.
Table 2. Clinical utility of the implementation of NGS for ALL patients.
Table 2. Clinical utility of the implementation of NGS for ALL patients.
MethodVariants DetectedGenetic SubtypesLimitations
WGSPoint mutations, aneuploidy, SV, CNVs, and BCR/TCR rearrangements. Useful for detecting germinal variants and polymorphism related to predisposition and treatment response, as well as GWASB-ALL: HeH, Hypodiploidy, BCR::ABL1, ETV6::RUNX1; TCF3::PBX1, KMT2Ar,
DUX4r; MEF2Dr; ZNF384r, NUTM1r; HLFr, BCL2::MYC; PAX5 (P80R); IKZF1 (N159Y); other related CNVs.
1. A considerable number of unknown significance or likely pathogenic status
2. Currently, this is considered a high-cost method based on sequencing-platform acquisition, data storage, and analysis.
T-ALL: BCL11B; TAL/LMO rearrangements, HOXA; SPI1; NKX2-1; TAL1 mutations
WESPoint mutations, SV, CNVs, and aneuploidies. Useful for detecting novel fusionsB-ALL: PAX5 (P80R), IKZF1 (N159Y), and mutations in the Ph-like group (ABL1, JAK)1. Detection of phenocopies
2. Limitation of the analysis to the detection of variants in coding regions
3. Less coverage (less sensitive)
Target Sequencing (DNA/RNA)Point mutations, aneuploidy, SV (InDels and gene fusions), and CNVs. Useful for increasing sensitivity (greater depth of coverage compared with WGS studies) for the detection of variants with low VAF; as well as variants of intron- and splicing-related regions variantsDefined target variants, such as BCR::ABL1, ETV6::RUNX1; TCF3::PBX1, BCL2::MYC; intrachromosomal amplifications (iAMP21, PAX5AMP), and other variants, such as CNVs (IKZF1, CDKN2A/2B, BTG1, PAX5); or mutations including PAX5 (P80R); or IKZF1 (N159Y)Non-targeted
alterations
WTSPoint mutations, CNVs, gene fusion, GEP, alternative splicing analysis, and BCR/TCR rearrangementsSubtypes defined by distinct gene expression profiles (phenocopies). Target variants such as BCR::ABL1, ETV6::RUNX1, TCF3::PBX1, and BCL2::MYC, novel fusions, PAX5 alteration and othersPloidy alterations and NOS T-ALL or B-other ALL
T-ALL: SET::NUP214, PICALM::MLLT10, NUP98::RAP1GDS1, TAL/LMO rearrangements
OGM *SV, aneuploidies, CNV, and balanced shifts in positionB-ALL: HeH, hypodiploidy, BCR-ABL1, ETV6::RUNX1; TCF3::PBX1, KMT2Ar, DUX4r, MEF2Dr; ZNF384r, NUTM1r, HLFr, BCL2::MYC; and other novel fusions1. Point mutation detection
2. Only fresh or preserved samples that guarantee adequate uHMW DNA integrity can be used.
T-ALL: TAL/LMO rearrangements KMT2A::PRDM10 and other novel fusions
Abbreviations. WGS: whole-genome sequencing, WES: whole-exome sequencing, WTS: whole-transcriptome sequencing, OGM: Optical Genome Mapping (GWAS: genome-wide association study, SV: structural variants, CNVs: copy number variants), HeH: high hyperdiploidy, BCR: B-cell receptor; TCR: T-cell receptor; uHMW: ultra-high molecular weight, NOS: not otherwise specified. OGM* is not an NGS technique; it does not determine the nucleotide sequence, but it allows for the analysis of the genome’s structure and organization.
Table 3. Comparison between different leukemia classifications based on genomic profiles.
Table 3. Comparison between different leukemia classifications based on genomic profiles.
WHO-HAEM5,
2022 Classification [55]
MLL, GMALL, and EGIL,
2022 Classification [57]
International Consensus
Classification (ICC), 2022 [58]
B-acute lymphoblastic leukemia (B-ALL)
Aneuploidies
High hyperdiploidyHigh hyperdiploidyHyperdiploidy
HypodiploidyHypodiploidyLow hypodiploidy
Near haploid
Gene fusion
BCR::ABL1 fusiont(9;22)(q34.1;q11.2)/BCR::ABL1t(9;22)(q34.1;q11.2)/BCR::ABL1
with lymphoid only involvement
with multilineage involvement
ETV6::RUNX1 fusiont(12;21)(p13.2;q22.1)/ETV6::RUNX1t(12;21)(p13.2;q22.1)/ETV6::RUNX1
IGH::IL3 fusiont(5;14)(q31.1;q32.3)/IL3::IGH t(5;14)(q31.1;q32.3)/IL3::IGH
TCF3::PBX1 fusiont(1;19)(q23.3;p13.3)/TCF3::PBX1t(1;19)(q23.3;p13.3)/TCF3::PBX1
TCF3::HLF fusion TCF3::HLF fusion
BCL2/MYC
Phenocopies
BCR::ABL1-like featuresBCR::ABL1–likeBCR::ABL1–like, ABL-1 class rearranged
BCR::ABL1–like, JAK-STAT activated
BCR::ABL1–like, NOS
ETV6::RUNX1-like features ETV6::RUNX1-likeETV6::RUNX1-like (provisional entity)
ZNF384 rearranged-like (provisional entity)
KMT2A rearranged-like (provisional entity)
Intrachromosomal amplifications
iAMP21iAMP21 iAMP21
Other specified rearrangements
KMT2A rearrangementt(v;11q23.3)/KMT2A rearranged t(v;11q23.3)/KMT2A rearranged
IG::MYC fusions MYC rearrangement
B-OtherDUX4 rearrangement DUX4 rearrangement
MEF2D rearrangementMEF2D rearrangement
ZNF384(362) rearrangement ZNF384(362) rearrangement
NUTM1 rearrangement NUTM1 rearrangement
HLF rearrangement
UBTF::ATXN7L3/PAN3,CDX2 (“CDX2/UBTF”)
Non Ph-like CRLF2 rearrangement
Point mutations
PAX5 P80R Mutated PAX5 (P80R) (provisional entity)
Mutated IKZF1 (N159Y)
Mutated ZEB2 (p.H1038R)/IGH::CEBPE (provisional entity)
Other B-ALL subtypes
PAX5 alterationPAX5 alteration (provisional entity)
B-ALL, NOSB-ALL, NOS
T-acute lymphoblastic leukemia (T-ALL)
Early T-cell precursor ALL, NOSEarly T-cell precursor ALL, NOSEarly T-cell precursor ALL, BCL11B-activated
T-ALL, NOST-ALL, NOSEarly T-cell precursor ALL, NOS
T-ALL, NOS
WHO-HAEM5, 2022 Classification. World Health Organization. B-lymphoblastic leukemia/lymphoma classification. MLL. Münchner Leukämielabor. GMALL. German Multicenter Study group for Adult Acute lymphoblastic leukemia. EGIL. European Group for Immunological Classification. NOS: not otherwise specified (when diagnostic criteria based on cytogenetic and molecular testing are applied, but subtyping is not possible even with exhaustive testing).
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

Ramírez Maldonado, V.; Navas Acosta, J.; Maldonado Marcos, I.; Villaverde Ramiro, Á.; Hernández-Sánchez, A.; Hernández Rivas, J.M.; Benito Sánchez, R. Unraveling the Genetic Heterogeneity of Acute Lymphoblastic Leukemia Based on NGS Applications. Cancers 2024, 16, 3965. https://doi.org/10.3390/cancers16233965

AMA Style

Ramírez Maldonado V, Navas Acosta J, Maldonado Marcos I, Villaverde Ramiro Á, Hernández-Sánchez A, Hernández Rivas JM, Benito Sánchez R. Unraveling the Genetic Heterogeneity of Acute Lymphoblastic Leukemia Based on NGS Applications. Cancers. 2024; 16(23):3965. https://doi.org/10.3390/cancers16233965

Chicago/Turabian Style

Ramírez Maldonado, Valentina, Josgrey Navas Acosta, Iván Maldonado Marcos, Ángela Villaverde Ramiro, Alberto Hernández-Sánchez, Jesús M. Hernández Rivas, and Rocío Benito Sánchez. 2024. "Unraveling the Genetic Heterogeneity of Acute Lymphoblastic Leukemia Based on NGS Applications" Cancers 16, no. 23: 3965. https://doi.org/10.3390/cancers16233965

APA Style

Ramírez Maldonado, V., Navas Acosta, J., Maldonado Marcos, I., Villaverde Ramiro, Á., Hernández-Sánchez, A., Hernández Rivas, J. M., & Benito Sánchez, R. (2024). Unraveling the Genetic Heterogeneity of Acute Lymphoblastic Leukemia Based on NGS Applications. Cancers, 16(23), 3965. https://doi.org/10.3390/cancers16233965

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

Article Metrics

Back to TopTop