Skip to main content

Spectrum and clinical features of gene mutations in Chinese pediatric acute lymphoblastic leukemia

Abstract

Purpose

The 5-year survival rate of children with acute lymphoblastic leukemia (ALL) is 85–90%, with a 10–15% rate of treatment failure. Next-generation sequencing (NGS) identified recurrent mutated genes in ALL that might alter the diagnosis, classification, prognostic stratification, treatment, and response to ALL. Few studies on gene mutations in Chinese pediatric ALL have been identified. Thus, an in-depth understanding of the biological characteristics of these patients is essential. The present study aimed to characterize the spectrum and clinical features of recurrent driver gene mutations in a single-center cohort of Chinese pediatric ALL.

Methods

We enrolled 219 patients with pediatric ALL in our single center. Targeted sequencing based on NGS was used to detect gene mutations in patients. The correlation was analyzed between gene mutation and clinical features, including patient characteristics, cytogenetics, genetic subtypes, risk stratification and treatment outcomes using χ2-square test or Fisher’s exact test for categorical variables.

Results

A total of 381 gene mutations were identified in 66 different genes in 152/219 patients. PIK3R1 mutation was more common in infants (P = 0.021). KRAS and FLT3 mutations were both more enriched in patients with hyperdiploidy (both P < 0.001). NRAS, PTPN11, FLT3, and KMT2D mutations were more common in patients who did not carry the fusion genes (all P < 0.050). PTEN mutation was significantly associated with high-risk ALL patients (P = 0.011), while NOTCH1 mutation was common in middle-risk ALL patients (P = 0.039). Patients with ETV6 or PHF6 mutations were less sensitive to steroid treatment (P = 0.033, P = 0.048, respectively).

Conclusion

This study depicted the specific genomic landscape of Chinese pediatric ALL and revealed the relevance between mutational spectrum and clinical features of Chinese pediatric ALL, which highlights the need for molecular classification, risk stratification, and prognosis evaluation.

Peer Review reports

Introduction

Pediatric acute lymphoblastic leukemia (ALL) is the most common childhood malignancy with 5-year overall survival (OS) rate of 85–90% and treatment failure rate of 10–15% [1,2,3,4]. Next-generation sequencing (NGS) identified recurrent mutated genes in pediatric ALL that might alter the diagnosis, classification, prognostic stratification, treatment, and response to ALL. However, there are few studies on gene mutations in Chinese pediatric ALL. Thus, an in-depth understanding of the biological characteristics of these patients is essential, and it is necessary to conduct comprehensive and thorough gene mutation detection by NGS in ALL patients. In the current study, we aimed to characterize the spectrum and clinical features of gene pathogenic variants in a single-center cohort of Chinese pediatric ALL patients.

Methods

Patients

The cohort in the present study consisted of 219 consecutive children (0.05–16.25, median: 3.75 years) with newly diagnosed ALL (n = 196, B-cell ALL (B-ALL); n = 23, T-cell ALL (T-ALL)) receiving treatment at our hospital between October 2017 and October 2019 and for whom complete data were available. The protocol was approved by the Medical Ethics Committee of the Children’s Hospital, and written informed consent was obtained from the parents or guardians. ALL was diagnosed based on the morphology, immunophenotyping, cytogenetics, and the molecular biology of leukemia cells [5]. Flow cytometric (FCM) immunophenotyping of bone marrow was performed on FACSCalibur with CellQuest software. A total of 51 fusion genes, including ETV6-RUNX1, BCR-ABL1, and MLL rearrangement, were examined by polymerase chain reaction (PCR). Karyotyping analysis was conducted by conventional methods. The patients were treated according to the National Protocol of Childhood Leukemia in China (NPCLC)-ALL2008 protocol, a modified form of protocol NPCAC97 [6].

Risk stratification

In this study, patients were classified into three groups: low risk (LR), intermediate risk (IR), and high risk (HR). If any one of the following criteria were fulfilled, the patients were assigned to the HR group: (i) Patients < 1-year-old; (ii) MLL rearrangement or t (17; 19) [TCF3/HLF] or hypodiploidy (< 45 chromosomes); (iii) Altered IKZF1; (iv) Poor response to prednisone or did not reach complete remission (CR) at the end of induction therapy; (v) Minimal residual disease (MRD) ≥ 1% on day 15 or 33 of induction. For IR group, any of the following criteria need to be met: (i) Patients > 10-year-old; (ii) White blood cell (WBC) count > 50 × 109 /L; (iii) T-ALL; (iv) TCF3/PBX1; (v) Ph+ ALL; (vi) Ph-like ALL; (vii) Central nervous system (CNS) 2 status (< 5 leukocytes/μL with blast cells in a cerebrospinal fluid sample with < 10 erythrocytes/μL), CNS3 status (≥ 5 leukocytes/μL with blast cells in a cerebrospinal fluid sample with < 10 erythrocytes/μL or the presence of a cerebral mass or cranial palsy) [7]; (viii) Testicular leukemia at diagnosis; (ix) 0.1% ≤ MRD ≤ 1% on day 15 of induction or 0.01% ≤ MRD ≤ 1% on day 33 of induction. The remaining patients with MRD < 0.01% on day 15 and 33 were categorized into the LR group.

Minimal residual disease monitoring

MRD was determined by FCM (FACSCalibur), as described previously [8]. The antibodies used for staining were conjugated with any of the following fluorochromes: fluorescein isothiocyanate (FITC), phycoerythrin (PE), peridinin chlorophyll protein (PerCP), and allophycocyanin (APC). The appropriate combination of antibodies was selected for MRD detection, such as leukemia-associated immunophenotype (LAIP). For B-ALL, CD19/CD10/CD34/CD45 was the most common combination; meanwhile, CD19/CD10/CD45/CD20, CD19/CD20/CD34/CD45, and CD19/CD10/CD34/CD20 were also utilized. For T-ALL, CD34, CD7, CD3, terminal deoxynucleotide transferase (TdT), and HLA-DR were the primary markers, and the most frequently used antibody combinations were CD45/CD2/CD3/CD56 and CD45/CD7/CD3/CD56. The bone marrow MRD was detected on days 15 and 33 after the start of induction.

Next-generation sequencing and mutation analysis

Genomic DNA was extracted from bone marrow samples at diagnosis. The spectrum of gene mutations was determined through NGS platform at Acornmed Biotechnology Co. Ltd. (Beijing, China). Genetic profiling included the targeted sequencing of 185 genes. Multiplex libraries were sequenced using Illumina NovaSeq. The following criteria were used to filter raw variant results: average effective sequencing depth on target per sample ≥ 1000x; variant allele frequency (VAF) ≥1% for single nucleotide variations (SNVs), insertions, or deletions (InDels); mapping quality ≥30; and base quality ≥30. The reads were aligned to the human genome using Burrows-Wheeler alignment (BWA, version 0.7.12). PCR duplicates were removed using the MarkDuplicates tool in Picard. Genome Analysis Toolkit (GATK; version 3.8) comprising of IndelRealigner and BaseRecalibrator was applied for realignment and recalibration of the BWA data, respectively. Mutect2 was used to identify SNVs and InDels. All the variants were annotated by ANNOVAR software, including 1000G projects, COSMIC, SIFT, and PolyPhen.

Statistics

The correlations between various gene mutations and clinical features were analyzed using χ2-square test or Fisher’s exact test for categorical variables A two-sided P value < 0.05 was considered to indicate statistical significance.

Results

Spectrum of gene mutations in pediatric ALL patients

The baseline characteristics of the patients are listed Table 1. A total of 381 gene mutations, including SNVs and InDels, were identified in 66 different genes in 152/219 patients (69%) (Fig. 1). The average number of gene mutations was 2.5 (range 0–10)/patient, and the average number of mutations in T-ALL and B-ALL patients was 3 (0–10) and 1 (0–7), respectively. The mutated genes were comprised of several functional groups, mainly including 54.4% in signaling pathway, 12% in transcription factors, and 11% in DNA methylation (Fig. 2). The mutated genes with the gene mutation fraction of patients ≥5% in our cohort were NRAS (n = 51, 23%), KRAS (n = 36, 16%), FLT3 (n = 23, 11%), PTPN11 (n = 15, 7%), NOTCH1 (n = 14, 6%) and KMT2D (n = 11, 5%), respectively (Fig. 1).

Table 1 Clinical features
Fig. 1
figure 1

The landscape of gene mutations identified by next-generation sequencing (NGS) in 219 Chinese pediatric patients with acute lymphoblastic leukemia (ALL)

Fig. 2
figure 2

Functional pathways of mutated genes. a Heatmap representing mutated genes classified into different functional pathways. b Proportion of gene function groups

The median VAF of gene mutations in the 219 ALL patients was 20% (Fig. 3). The mutated genes with median VAF ≥ 25% included CBL, TET2, CDKM2A, BCORL1, EZH2, TNFAIP3, FAT1, BRAF, etc. Conversely, the rest of the mutated genes with lower VAF < 25% indicated that they were present in a subpopulation of the sequenced cells. Notably, four RAS signaling pathway mutated genes (FLT3, NRAS, KRAS, and PTPN11) had lower VAFs < 25%, suggesting that these genes were subclones. The mutations mainly affected codons 12 and 13 of KRAS and NRAS. The allele frequency of different hotspot mutations was compared, and it was found that the frequency of KRAS G12D was higher than that of other hotspot mutations (Supplemental Fig. S1), which was similar to the results reported in some previous literatures [9].

Fig. 3
figure 3

Variant allele frequency (VAF) analysis of various gene mutations

Next, we investigated the co-occurrence of mutated genes and found significant associations between mutated NOTCH1 and mutations in FBXW7 and PTEN, mutated JAK2 and mutations in MSH6 and PCLO, and mutated DNM2 and mutations in PHF6 and USP7. Moreover, pairwise associations were observed between ETV6 and KRAS, KDM6A and KMT2D, RUNX1 and ATRX, ASXL1 and SH2B3, and USP7 and FBXW7 (P < 0.05, Fig. 4).

Fig. 4
figure 4

Correlation analysis of pairwise co-occurring mutated genes

Correlation between gene mutations and patient characteristics, cytogenetics

The correlation analysis between gene mutations and the patient characteristics in these patients revealed that PIK3R1 mutation was more common in infants compared to patients ≥1-year-old (P = 0.021, Fig. 5A). Moreover, PIK3R1was more common in infants, compared with 1 ~ 10-year-old group (P < 0.001, Fig. 5B).While compared with infant group, both FLT3 and KRAS mutations were less common in patients > 10-year-old (P = 0.023, P = 0.006, respectively, Fig. 5C). At the same time, we discovered PIK3R1, PCLO, USP7, DNM2 and SETD2 mutations were more common in > 10-year-old group, compared with 1 ~ 10-year-old group (all P < 0.05, Fig. 5D). We also found that gender did not influence the mutational status of any of the genes. In addition, the mutations in NOTCH1 and PTEN were more common in patients with initial leukocyte count > 50 × 109/L (P < 0.001 and P = 0.097, respectively) (Fig. 5E). Patients with FLT3 mutations showed lower platelet counts (≤62 × 109/L, P = 0.024, Fig. 5F) and hemoglobin level (≤82 g/L, P = 0.049, Fig. 5G) at diagnosis than those without FLT3 mutations. While patients with NOTCH1 mutations had a high hemoglobin level (> 82 g/L, P = 0.005) (Fig. 5E). Compared to B-ALL, NOTCH1, PTEN, FBXW7, USP7, DNM2, and CDKN2A were more frequently mutated in T-ALL (all P < 0.05) (Fig. 5H). Futhermore, KRAS and FLT3 mutations were both enriched in patients with hyperdiploidy (both P < 0.001, Fig. 5I).

Fig. 5
figure 5

Correlation between gene mutations and patient characteristics and cytogenetics. a-d Association of gene mutations with different age groups of patients. (E-G) Association of gene mutations with white blood cell e, platelet f, and hemoglobin g groups, respectively. h Different distribution of gene mutations in T-ALL and B-ALL. i Association of gene mutations between hyperdiploidy and not hyperdiploidy groups

Correlation between gene mutations and genetic subtypes, risk stratification and treatment outcomes

Molecular genetic analyses of 51 fusion transcripts, including ETV6-RUNX1, TCF3-PBX1, BCR-ABL1, and KMT2A (MLL) rearrangement were conducted successfully in all the patients. Strikingly, NRAS, PTPN11, FLT3, and KMT2D mutations were common in patients who did not carry the fusion genes (all P < 0.050), and NRAS mutations were rarely in patients with ETV6-RUNX1 (P = 0.002). RUNX1 and ROBO1 mutations were more closely linked to BCR-ABL1 fusion gene (P = 0.017, P = 0.032, respectively), and PAX5, PHF6, and STAG2 mutations were associated with TCF3-PBX1 fusion gene (P = 0.001, P = 0.006, P = 0.046, respectively). MLL translocations co-existed with PIK3R1 mutation (P = 0.017, Fig. 6A). Some genes are closely associated with the prognosis of the disease. Herein, we studied the associations of gene mutations with risk stratification in the cohort. PTEN mutation was significantly associated with high risk ALL patients (P = 0.011), while NOTCH1 mutation was common in intermediate risk ALL patients (P = 0.039, Fig. 6B). In addition, PIK3R1 mutation occurs frequently in high risk B-ALL patients (P = 0.023, Fig. 6C). Patients with ETV6 or PHF6 mutations detected at the time of diagnosis were less sensitive to steroid treatment (P = 0.033, P = 0.048, respectively, Fig. 6D). In B-ALL patients, we analyzed the associations between gene mutations and early MRD levels (MRD1 on day 15 and MRD2 on day 33). Patients with mutated PIK3R1, TET2, and KMT2D had a markedly higher MRD1 level (MRD1 ≥ 10− 2). ASXL1 and NRAS mutations were enriched in patients with 10− 3 ≤ MRD1 < 10− 2 (P = 0.024, P < 0.001, respectively), whereas NRAS and CREBBP mutations were less significantly in B-ALL patients with MRD1 < 10− 3 (P = 0.005, P = 0.042, Fig. 6E). However, no significant association was established between gene mutations and MRD2, but the patients carrying the genes with mutated chromatin modifiers exhibited a significantly high level of MRD2 (MRD ≥10− 4, 21.4% vs. 4%, P = 0.029, Fig. 6F).

Fig. 6
figure 6

Correlation between gene mutations and genetic subtypes, risk stratification and treatment outcomes. a Association of gene mutations with genetic fusion subtypes. b Association of gene mutations with risk stratification in all ALL patients. c Association of gene mutations with risk stratification in B-ALL. d Association of gene mutations with steroid therapeutic effects. e Association of gene mutations with MRD level in B-ALL. f Association between gene groups and MRD2 in B-ALL

Discussion

In this study, we dissected the genetic landscape, analyzed the mutational spectrum of various immunological ALL lineages, and explored the correlations between mutational and clinical features, including patient characteristics, risk stratification, and treatment outcomes in a Chinese pediatric ALL cohort. A number of gene pathogenic variants were identified, which provided a comprehensive genomic profile of Chinese pediatric ALL. Consistent with previous reports, B-ALL and T-ALL presented a distinct mutation spectrum; Ras pathway mutations were enriched in B-ALL, while Notch pathway mutations were enriched in T-ALL [10,11,12,13].

As described previously, mutations involved in the Ras signaling pathway (NRAS, KRAS, FLT3, PTPN11, and NF1) occurred in more than half of B-ALL patients [14, 15] . Also, a higher incidence of mutations was detected in NRAS rather than KRAS. This finding was contradictory to the previous studies in the Chinese cohort [15] but was in agreement to that from the USA, Sweden, and Korea [14,15,16,17,18,19]. These discrepancies might be related to the population distribution and environmental factors, which highlighted the genetic heterogeneity of pediatric ALL. Compared to CBL, TET2, CDKM2A, and BCORL1 genes with a higher median VAF, Ras signaling pathway-related genes, such as FLT3, NRAS, KRAS, and PTPN11, displayed a lower median VAF of 5–20%. The lower VAF indicated that Ras mutations were more likely subclones rather than a major clone [20], suggesting that B-ALL is driven by other fusion genes. Reportedly, Ras pathway functioned as a molecular switch for signaling pathways that regulated cell proliferation, survival, growth, migration, and differentiation [21]. Thus, we speculated that Ras pathway mutations occurred during B-ALL progression rather than tumorigenesis. Based on genetic testing of a large number of ALL patients, Shu et al. and Perentesis et al. demonstrated that RAS mutations did not present any unique clinical manifestation nor predicted clinical outcomes [18, 19]. Moreover, some recent studies showed that ALL patients with Ras pathway mutations, especially KRAS/NRAS mutations, present high-risk features, including early relapse and CNS involvement [22,23,24]. In our cohort, no correlation was established between the presence of Ras mutation and clinical characteristics, risk stratification, and MRD level. This phenomenon could be attributed to the neutralization effect of other genomic variations, such as low-risk hyperdiploidy and high-risk hypodiploidy on prognosis [25,26,27,28]. To determine whether Ras pathway status influences the clinical characteristics and risk stratification, additional studies are warranted on various cytogenetic subgroups of B-ALL. Notch pathway mutations, especially NOTCH1 and FBXW7, were enriched in T-ALL patients, as previously reported [29,30,31,32]. NOTCH1 was the most common mutated gene in about 60.9% of all T-ALL cases, followed by PTEN (21.7%) and FBXW7 (21.7%). Notch signaling pathway, especially NOTCH1, plays a crucial role in all stages of T lymphocyte development and can promote the differentiation of lymphoid precursor cells into T lymphocytes and inhibit their differentiation into B lymphocytes [33, 34]. Except for the excessive activation of the Notch pathway, impaired CDKN2A/2B cell cycle regulators also played a prominent role in T-ALL pathogenesis. Strikingly, CDKN2A/2B deletions were detected in > 50% of T-ALL cases [10, 35]. However, the copy number variations were not detected and analyzed in the present study, and only a few CDKN2A gene point mutations were identified in T-ALL. Recent sequencing studies demonstrated that T-ALL was an aggressive malignancy caused by the accumulation of genomic lesions. On average, 10–20 mutations were detected in T-ALL cells [10, 36,37,38]. Although our study showed that T-ALL had a significantly higher mutation level than B-ALL, the average number of mutations was still lower than the expected value. This deviation could be attributed to the scope of sequencing, the evaluated variation types, the sensitivity of the test, and the filter criteria of mutation calling. We also found that the accumulation of mutations in T-ALL did not occur randomly [39]. Interestingly, the coexistence of NOTCH1-PTEN-FBXW7 and DNM2-USP7-PHF6 mutations was observed in our T-ALL cohort. The coexistence phenomenon suggested that those Notch pathway and non-Notch pathway genes interconnect physiologically and cooperate during the development and progression of the T-ALL, respectively. MLL translocations and PIK3R1 mutations were common in infant ALL, a group characterized as immature cytologically, resistant to conventional therapies, and showing poor prognosis. Moreover, a significant coexistence between MLL gene arrangement and PIK3R1 mutations was detected in our cohort. This observation indicated that PI3K/AKT is a secondary hit for partial MLL-positive ALL.

In this study, we wanted to analyze the relationship between these pathogenic variants and prognosis. Due to the short follow-up time, we only analyzed the relationship between MRD of day 15, 33 of induction and mutations. MRD is showed a high level in patients with PIK3R1, TET2, and KMT2D mutations, indicating a high risk of relapse. Both TET2 and KMT2D belong to epigenetic regulator genes, which play key roles in DNA demethylation and histone H3 methylation, respectively [13, 40]. This finding suggested that mutations in epigenetic regulator genes elevate the MRD level. Future studies we will continue to explore the relationships between long-term prognosis, recurrence rates and pathogenic variants, which may help improve the prognosis of ALL in children.

In summary, our study depicted the specific genomic landscape and revealed the relevance between mutational spectrum and clinical features of Chinese pediatric ALL in a single cohort, including patient characteristics, cytogenetics, genetic subtypes, risk stratification and treatment outcomes. The discovery of this mutational spectrum highlights the need for molecular classification, risk stratification, and prognosis evaluation and also provide the basis for the development and application of new targeted therapy for pediatric ALL.

Availability of data and materials

Data and materials are available in supplementary file 1.

Abbreviations

NGS:

Next-generation sequencing

ALL:

acute lymphoblastic leukemia

OS:

overall survival

FCM:

flow cytometric

PCR:

polymerase chain reaction

LR:

low risk

IR:

intermediate risk

HR:

high risk

CR:

complete remission

MRD:

minimal residual disease

WBC:

white blood cell

CNS:

central nervous system

FITC:

fluorescein isothiocyanate

PE:

phycoerythrin

PerCP:

peridinin chlorophyll protein

APC:

allophycocyanin

LAIP:

leukemia-associated immunophenotype

VAF:

variant allele frequency

SNVs:

single nucleotide variations

BWA:

Burrows-Wheeler alignment

GATK:

Genome Analysis Toolkit

InDels:

deletions

B-ALL:

B-cell ALL

T-ALL:

T-cell ALL

TdT:

terminal deoxynucleotide transferase

References

  1. Dores GM, Devesa SS, Curtis RE, Linet MS, Morton LM. Acute leukemia incidence and patient survival among children and adults in the United States, 2001-2007. Blood. 2012;119(1):34–43. https://doi.org/10.1182/blood-2011-04-347872.

    Article  CAS  Google Scholar 

  2. Inaba H, Greaves M, Mullighan CG. Acute lymphoblastic leukaemia. Lancet. 2013;381(9881):1943–55. https://doi.org/10.1016/S0140-6736(12)62187-4.

    Article  Google Scholar 

  3. Liao C, et al. Minimal residual disease-guided risk Restratification and therapy improves the survival of childhood acute lymphoblastic leukemia: experience from a tertiary Children's Hospital in China. J Pediatr Hematol Oncol. 2019;41(6):e346–54. https://doi.org/10.1097/MPH.0000000000001412.

    Article  Google Scholar 

  4. Sasaki K, et al. 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(6):650–8. https://doi.org/10.1002/ajh.26156.

    Article  CAS  Google Scholar 

  5. Sabattini E, Bacci F, Sagramoso C, Pileri SA. WHO classification of tumours of haematopoietic and lymphoid tissues in 2008: an overview. Pathologica. 2010;102(3):83–7.

    CAS  Google Scholar 

  6. Tang Y, et al. Long-term outcome of childhood acute lymphoblastic leukemia treated in China. Pediatr Blood Cancer. 2008;51(3):380–6. https://doi.org/10.1002/pbc.21629.

    Article  Google Scholar 

  7. Pui CH, et al. Treating childhood acute lymphoblastic leukemia without cranial irradiation. N Engl J Med. 2009;360(26):2730–41. https://doi.org/10.1056/NEJMoa0900386.

    Article  CAS  Google Scholar 

  8. Xu XJ, et al. Day 22 of induction therapy is important for minimal residual disease assessment by flow cytometry in childhood acute lymphoblastic leukemia. Leuk Res. 2012;36(8):1022–7. https://doi.org/10.1016/j.leukres.2012.03.014.

    Article  Google Scholar 

  9. Antić Ž, et al. Multiclonal complexity of pediatric acute lymphoblastic leukemia and the prognostic relevance of subclonal mutations. Haematologica. 2021;106(12):3046–55. https://doi.org/10.3324/haematol.2020.259226.

    Article  CAS  Google Scholar 

  10. Girardi T, Vicente C, Cools J, De Keersmaecker K. The genetics and molecular biology of T-ALL. Blood. 2017;129(9):1113–23. https://doi.org/10.1182/blood-2016-10-706465.

    Article  CAS  Google Scholar 

  11. Roberts KG, Brady SW, Gu Z, Shi L, Mullighan CG. The genomic landscape of childhood acute lymphoblastic leukemia. BLOOD. 2019;134:649–9.

    Article  Google Scholar 

  12. Tasian SK, Hunger SP. Genomic characterization of paediatric acute lymphoblastic leukaemia: an opportunity for precision medicine therapeutics. Br J Haematol. 2017;176(6):867–82. https://doi.org/10.1111/bjh.14474.

    Article  CAS  Google Scholar 

  13. Zhang H, et al. Genetic mutational analysis of pediatric acute lymphoblastic leukemia from a single center in China using exon sequencing. BMC Cancer. 2020;20(1):211. https://doi.org/10.1186/s12885-020-6709-7.

    Article  CAS  Google Scholar 

  14. Yoo JW, et al. Spectrum of genetic mutations in Korean pediatric acute lymphoblastic leukemia. J Clin Med. 2022;11(21). https://doi.org/10.3390/jcm11216298.

  15. Liu Y, et al. Genomic and clinical analysis of children with acute lymphoblastic leukemia. Comput Math Methods Med. 2022;2022:7904293. https://doi.org/10.1155/2022/7904293.

    Article  Google Scholar 

  16. Al-Kzayer LF, et al. Analysis of KRAS and NRAS gene mutations in Arab Asian children with acute leukemia: high frequency of RAS mutations in acute lymphoblastic leukemia. Pediatr Blood Cancer. 2015;62(12):2157–61. https://doi.org/10.1002/pbc.25683.

    Article  CAS  Google Scholar 

  17. Paulsson K, et al. Mutations of FLT3, NRAS, KRAS, and PTPN11 are frequent and possibly mutually exclusive in high hyperdiploid childhood acute lymphoblastic leukemia. Genes Chromosomes Cancer. 2008;47(1):26–33. https://doi.org/10.1002/gcc.20502.

    Article  CAS  Google Scholar 

  18. Perentesis JP, et al. RAS oncogene mutations and outcome of therapy for childhood acute lymphoblastic leukemia. Leukemia. 2004;18(4):685–92. https://doi.org/10.1038/sj.leu.2403272.

    Article  CAS  Google Scholar 

  19. Shu XO, et al. Parental exposure to medications and hydrocarbons and ras mutations in children with acute lymphoblastic leukemia: a report from the Children's oncology group. Cancer Epidemiol Biomark Prev. 2004;13(7):1230–5.

    Article  CAS  Google Scholar 

  20. Oshima K, et al. Mutational landscape, clonal evolution patterns, and role of RAS mutations in relapsed acute lymphoblastic leukemia. Proc Natl Acad Sci U S A. 2016;113(40):11306–11. https://doi.org/10.1073/pnas.1608420113.

    Article  CAS  Google Scholar 

  21. Zhang HH, et al. Ras pathway mutation feature in the same individuals at diagnosis and relapse of childhood acute lymphoblastic leukemia. Transl Pediatr. 2020;9(1):4–12. https://doi.org/10.21037/tp.2020.01.07.

    Article  Google Scholar 

  22. Ney GM, et al. Mutations predictive of hyperactive Ras signaling correlate with inferior survival across high-risk pediatric acute leukemia. Transl Pediatr. 2020;9(1):43–50. https://doi.org/10.21037/tp.2019.12.03.

    Article  Google Scholar 

  23. Reshmi SC, et al. Targetable kinase gene fusions in high-risk B-ALL: a study from the Children's oncology group. Blood. 2017;129(25):3352–61. https://doi.org/10.1182/blood-2016-12-758979.

    Article  CAS  Google Scholar 

  24. Takashima Y, et al. Target amplicon exome-sequencing identifies promising diagnosis and prognostic markers involved in RTK-RAS and PI3K-AKT signaling as central oncopathways in primary central nervous system lymphoma. Oncotarget. 2018;9(44):27471–86. https://doi.org/10.18632/oncotarget.25463.

    Article  Google Scholar 

  25. Davidsson J, et al. Relapsed childhood high hyperdiploid acute lymphoblastic leukemia: presence of preleukemic ancestral clones and the secondary nature of microdeletions and RTK-RAS mutations. Leukemia. 2010;24(5):924–31. https://doi.org/10.1038/leu.2010.39.

    Article  CAS  Google Scholar 

  26. Wiemels JL, et al. Backtracking RAS mutations in high hyperdiploid childhood acute lymphoblastic leukemia. Blood Cells Mol Dis. 2010;45(3):186–91. https://doi.org/10.1016/j.bcmd.2010.07.007.

    Article  CAS  Google Scholar 

  27. Zhang J, et al. Key pathways are frequently mutated in high-risk childhood acute lymphoblastic leukemia: a report from the Children's oncology group. Blood. 2011;118(11):3080–7. https://doi.org/10.1182/blood-2011-03-341412.

    Article  CAS  Google Scholar 

  28. Irving JA, et al. Integration of genetic and clinical risk factors improves prognostication in relapsed childhood B-cell precursor acute lymphoblastic leukemia. Blood. 2016;128(7):911–22. https://doi.org/10.1182/blood-2016-03-704973.

    Article  CAS  Google Scholar 

  29. Eguchi-Ishimae M, Eguchi M, Kempski H, Greaves M. NOTCH1 mutation can be an early, prenatal genetic event in T-ALL. Blood. 2008;111:376–8. https://doi.org/10.1182/blood-2007-02-074690.

    Article  CAS  Google Scholar 

  30. Erbilgin Y, et al. Prognostic significance of NOTCH1 and FBXW7 mutations in pediatric T-ALL. Dis Markers. 2010;28(6):353–60. https://doi.org/10.3233/DMA-2010-0715.

    Article  CAS  Google Scholar 

  31. Sulis ML, et al. A novel class of activating mutations in NOTCH1 in T-ALL. BLOOD. 2007.

  32. Valliyammai N, Nancy NK, Sagar TG, Rajkumar T. Study of NOTCH1 and FBXW7 mutations and its prognostic significance in south Indian T-cell acute lymphoblastic leukemia. J Pediatr Hematol Oncol. 2018;40(1):e1–8. https://doi.org/10.1097/MPH.0000000000001006.

    Article  CAS  Google Scholar 

  33. Radtke F, MacDonald HR, Tacchini-Cottier F. Regulation of innate and adaptive immunity by notch. Nat Rev Immunol. 2013;13(6):427–37. https://doi.org/10.1038/nri3445.

    Article  CAS  Google Scholar 

  34. Vanderbeck A, Maillard I. Notch signaling at the crossroads of innate and adaptive immunity. J Leukoc Biol. 2020;109:535–48. https://doi.org/10.1002/JLB.1RI0520-138R.

    Article  CAS  Google Scholar 

  35. Yeh TC, et al. Clinical and biological relevance of genetic alterations in pediatric T-cell acute lymphoblastic leukemia in Taiwan. Pediatr Blood Cancer. 2019;66(1):e27496. https://doi.org/10.1002/pbc.27496.

    Article  CAS  Google Scholar 

  36. De Keersmaecker K, et al. Exome sequencing identifies mutation in CNOT3 and ribosomal genes RPL5 and RPL10 in T-cell acute lymphoblastic leukemia. Nat Genet. 2013;45(2):186–90. https://doi.org/10.1038/ng.2508.

    Article  CAS  Google Scholar 

  37. Holmfeldt L, et al. The genomic landscape of hypodiploid acute lymphoblastic leukemia. Nat Genet. 2013;45(3):242–52. https://doi.org/10.1038/ng.2532.

    Article  CAS  Google Scholar 

  38. Zhang J, et al. The genetic basis of early T-cell precursor acute lymphoblastic leukaemia. Nature. 2012;481(7380):157–63. https://doi.org/10.1038/nature10725.

    Article  CAS  Google Scholar 

  39. Vogelstein B, et al. Cancer genome landscapes. Science. 2013;339(6127):1546–58. https://doi.org/10.1126/science.1235122.

    Article  CAS  Google Scholar 

  40. Ichiyama K, et al. The methylcytosine dioxygenase Tet2 promotes DNA demethylation and activation of cytokine gene expression in T cells. Immunity. 2015;42(4):613–26. https://doi.org/10.1016/j.immuni.2015.03.005.

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank all the staffs from both hematology-oncology devision and hematology-oncology laboratory.

Funding

This study was partly supported by grants from National Natural Science Foundation of China (No: 81770202, No.81470304) and Pediatric Leukemia Diagnosis and Therapeutic Technology Research Center of Zhejiang Province (JBZX-201904).

Author information

Authors and Affiliations

Authors

Contributions

Guarantor of integrity of entire study: Yongmin Tang, Diying Shen. Study concepts: Yongmin Tang, Xiaojun Xu, Hua Song. Study design: Diying Shen, Jingying Zhang, Weiqun Xu. Literature research: Diying Shen, Lixia Liu. Clinical studies: Fenying Zhao, Juan Liang, Chan Liao, Yan Wang, Tian Xia. Experimental studies: Chengcheng Wang, Feng Lou, Shanbo Cao. Data acquisition: Diying Shen. Data analysis/interpretation: Diying Shen, Lixia Liu. Statistical analysis: Diying Shen, Lixia Liu. Manuscript preparation: Diying Shen, Lixia Liu. Manuscript editing: Diying Shen, Jiayue Qin. Manuscript revision/review: Yongmin Tang. Manuscript final version approval: Diying Shen, Yongmin Tang.

Corresponding author

Correspondence to Yongmin Tang.

Ethics declarations

Ethics approval and consent to participate

The experimental protocol was established, according to the ethical guidelines of the Helsinki Declaration and was approved by the Human Ethics Committee of the Children’s Hospital, Zhejiang University School of Medicine (Ethical code:2021-IRB-054). Written informed consent was obtained from individual or guardian participants.

Consent for publication

Not applicable.

Competing interests

Authors declare no conflict of interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1:

 Figure S1. The variant allele frequency (VAF) of different KRAS and NRAS hotspot mutations.

Additional file 2:

 Table S2.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Shen, D., Liu, L., Xu, X. et al. Spectrum and clinical features of gene mutations in Chinese pediatric acute lymphoblastic leukemia. BMC Pediatr 23, 62 (2023). https://doi.org/10.1186/s12887-023-03856-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12887-023-03856-y

Keywords