Next Article in Journal
Large-Scale Biogeographical Shifts of Abundance of Antibiotic Resistance Genes and Marine Bacterial Communities as Their Carriers along a Trophic Gradient
Next Article in Special Issue
Comprehensive Molecular Profiling of NPM1-Mutated Acute Myeloid Leukemia Using RNAseq Approach
Previous Article in Journal
The Influence of Genetic Polymorphisms on the Expression of Interleukin-1beta, Prostaglandin E2 and Tumor Necrosis Factor Alpha in Peri-Implant Crevicular Fluid: A Cross-Sectional Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Germline Variants and Characteristic Features of Hereditary Hematological Malignancy Syndrome

by
Hironori Arai
1,2,
Hirotaka Matsui
3,4,
SungGi Chi
1,
Yoshikazu Utsu
2,
Shinichi Masuda
2,
Nobuyuki Aotsuka
2 and
Yosuke Minami
1,*
1
Department of Hematology, National Cancer Center Hospital East, Kashiwa 277-8577, Japan
2
Department of Hematology and Oncology, Japanese Red Cross Narita Hospital, Iidacho, Narita 286-0041, Japan
3
Department of Laboratory Medicine, National Cancer Center Hospital, Tsukiji, Chuoku 104-0045, Japan
4
Department of Medical Oncology and Translational Research, Graduate School of Medical Sciences, Kumamoto University, Kumamoto 860-8665, Japan
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(1), 652; https://doi.org/10.3390/ijms25010652
Submission received: 7 November 2023 / Revised: 25 December 2023 / Accepted: 29 December 2023 / Published: 4 January 2024

Abstract

:
Due to the proliferation of genetic testing, pathogenic germline variants predisposing to hereditary hematological malignancy syndrome (HHMS) have been identified in an increasing number of genes. Consequently, the field of HHMS is gaining recognition among clinicians and scientists worldwide. Patients with germline genetic abnormalities often have poor outcomes and are candidates for allogeneic hematopoietic stem cell transplantation (HSCT). However, HSCT using blood from a related donor should be carefully considered because of the risk that the patient may inherit a pathogenic variant. At present, we now face the challenge of incorporating these advances into clinical practice for patients with myelodysplastic syndrome (MDS) or acute myeloid leukemia (AML) and optimizing the management and surveillance of patients and asymptomatic carriers, with the limitation that evidence-based guidelines are often inadequate. The 2016 revision of the WHO classification added a new section on myeloid malignant neoplasms, including MDS and AML with germline predisposition. The main syndromes can be classified into three groups. Those without pre-existing disease or organ dysfunction; DDX41, TP53, CEBPA, those with pre-existing platelet disorders; ANKRD26, ETV6, RUNX1, and those with other organ dysfunctions; SAMD9/SAMD9L, GATA2, and inherited bone marrow failure syndromes. In this review, we will outline the role of the genes involved in HHMS in order to clarify our understanding of HHMS.

1. Introduction

Most hematologic malignancies are thought to spontaneously arise due to acquired genetic lesions in hematopoietic stem and precursor cells (HSPCs) [1]. However, in some cases of acute myeloid leukemia (AML) and myelodysplastic syndrome (MDS), a hereditary (mainly autosomal dominant) predisposition has been observed [2,3]. Typically, a family in which two or more first- or second-degree relatives have developed acute leukemia (AL), myeloid malignancies, characteristic cytopenias, or either MDS or AML, is defined as “familial MDS/AML”, or, more broadly, hereditary hematologic malignancy syndrome (HHMS) [4,5,6]. The field of HHMS has gained increasing recognition among clinicians and scientists worldwide. Both myeloid and lymphoid malignancies may be present in individuals or families with these syndromes. Genetic predisposition should be considered in patients who present with bone marrow failure, MDS, or AML at a young age or who present with unexpected hematologic toxicity during treatment for malignancy at a young age [7,8]. Identifying characteristics of such patients include physical abnormalities, endocrine abnormalities, short stature, stunted growth, and immunodeficiency in patients with hematologic abnormalities such as cytopenia, unexplained macro-erythroblastosis, or overt malignancy. A genetic MDS/AML predisposition may also be indicated by a family history of first- or second-degree relatives with malignancy, cytopenia, congenital abnormalities, or excessive toxicity from chemotherapy or radiation therapy [9]. However, the absence of characteristic clinical features or a negative family history does not exclude the presence of a germline MDS/AML syndrome. Germline variants may occur de novo or result from parental gonadal mosaicism [10]. HHMS often shows marked inter- and intra-familial differences in latency, phenotype, expression, and penetrance. For example, some germline MDS syndromes lack obvious syndromic features or have variable penetrance or delayed expression. Cytogenetic clonal abnormalities common to certain inherited MDS disorders may warrant further investigation [11]. MDS with monosomy 7 frequently occurs in patients with germline variants in GATA-binding factor 2 (GATA2), sterile alpha motif domain containing 9 (SAMD9), sterile alpha motif domain containing 9 like (SAMD9L), or hereditary bone marrow failure syndrome [12]. Moreover, the involvement of hematopoietic transcription factor genes, such as CCAAT enhancer binding protein alpha (CEBPA), GATA2, runt-related transcription factor 1 (RUNX1), ankyrin repeat domain containing 26 (ANKRD26), and ETS variant transcription factor 6 (ETV6), is traditionally associated with solid tumors such as MutS homolog 6 (MSH6) and breast cancer gene 1 (BRCA1). Moreover, the recently identified genes DEAD-box helicase 41 (DDX41), SAMD9, SAMD9L are involved in leukemogenesis [13,14,15]. Many are found to be non-symptomatic and occur in various age groups. Studies suggest that about 10% of children and adults with MDS or AML may have heritable variants [5]. Importantly, these germline genetic abnormalities are not exclusive to the patient and may be shared by blood relatives, necessitating the screening of blood relatives. As our diagnostic capabilities in HHMS improve, we now face the challenge of incorporating these advances into clinical practice with MDS/AML patients and learning how to optimize the management and surveillance of patients and asymptomatic carriers [16].
The discovery of novel syndromes combined with the clinical, genetic, and epigenetic profiling of tumor samples has highlighted unique patterns of disease progression in HHMS. Despite these advances, causative lesions are identified in fewer than half of familial cases, and evidence-based guidelines are often inadequate. In the 2016 revision of the WHO classification, a new section was added for myeloid neoplasms with a germline predisposition, including cases of MDS, myeloproliferative neoplasms (MPN), and ALs that develop on a background of predisposing germline variants [17]. As part of the diagnosis, specific underlying genetic abnormalities or predisposing syndromes should be considered. The major syndromes can be categorized into the following three groups: those without preexisting disease or organ dysfunction [e.g., DDX41, tumor protein p53 (TP53), and CEBPA], those with pre-existing platelet disorders [e.g., ANKRD26, ETV6, and RUNX1], and those with organ dysfunction [e.g., SAMD9/SAMD9L, GATA2, and inherited bone marrow failure syndromes (IBMFSs)]. This review will outline the genes involved in the above HHMS (Table 1).

2. Myeloid Neoplasms without a Preexisting Disorder or Organ Dysfunction

2.1. Myeloid Neoplasms with a Germline DDX41 Variant

RNA helicases are a series of enzymes that remodel RNA–RNA or RNA–protein interactions in an NTP-dependent manner. Humans have more than 70 helicases that are classified into superfamily (SF) 1 and SF2 based on differences in sequence motifs within the helicase core domain [59,60]. SF1 includes Upf1-like RNA helicases, while SF2 includes the DEAD-box, DEAH-box/RNA helicase A-like, Ski2-like, and RIG-I-like families, with the DEAD-box family RNA helicases being the most numerous. While the DEAH-box RNA helicases are thought to translocate along the substrate RNA for remodeling, DEAD-box RNA helicases unwind substrate RNA locally; the mechanism of action of each is thus different, but they both play roles in virtually all processes that require RNA conformational changes, such as RNA transport, translation, RNA degradation, RNA splicing, and ribosome synthesis. As a single RNA helicase often exerts enzymatic activity in multiple cellular processes, it remains difficult to fully elucidate the pathogenesis of diseases due to abnormalities in RNA helicases.
In myeloid neoplasms, pathogenic variants in the gene encoding DDX41, a DEAD-box RNA helicase, are found in about 5% of cases [61]. It was recently shown that up to 13% of myeloid neoplasms have a genetic background [62], of which DDX41 variants account for about 80% of cases. MDS and AML occur in individuals with a heterozygous germline frameshift variant or a missense variant within the DEAD-box domain of DDX41 by later acquiring a somatic variant in the other allele, typically p.R525H (or p.G530D, etc., in a few cases) within the helicase domain [61,63,64] (Figure 1A). While many myeloid neoplasms with a genetic background develop at younger ages than those without a known genetic background, myeloid neoplasms with DDX41 variants are characterized by a late disease onset (mean age, 65 years) [22,23], which may hinder the identification of this gene as one of the genes responsible for genetic predisposition for myeloid leukemogenesis. In addition, the disease with a DDX41 variant is characterized by male dominancy, fewer proliferating tumor cells, hypoplastic bone marrow, and unique co-existing gene mutational patterns as compared to those in other myeloid neoplasms [65,66], with only DDX41 variants being identified in many cases [61], suggesting a unique disease pathogenesis of myeloid neoplasms with DDX41 variants. In contrast, the disease phenotype may differ between cases with a single DDX41 variant and biallelic variants [67], and a report suggest that there is no clear difference in disease phenotype between cases with known pathogenic DDX41 variants and variants of unknown significance (VUS) [68]. Consequently, it is necessary to establish a validation system and database that can accurately interpret the significance of individual variants.
  • A combination of germline and somatic DDX41 variants confers myeloid disease development.
Hematopoietic cells with a germline DDX41 variant acquire a somatic DDX41 variant at an advanced age. Myeloid neoplasms are thought to develop shortly after biallelic DDX41 variant acquisition, with or without the addition of a limited number of somatic variants in DNA repair-related genes, including CUX1 and TP53. It is also suggested that minor clones with biallelic DDX41 variants affect hematopoiesis by interfering with other cells [37].
B.
R-loop formation and its consequence.
R-loop accumulation due to impaired RNA splicing or other causes increases DNA replication stress and innate immune response, resulting in deficient hematopoiesis and leukemogenesis.
The prognosis of myeloid neoplasms with DDX41 variants is not necessarily worse than for those without a known genetic background, regardless of the tendency to be categorized as high-risk. However, the development of disease at advanced ages often makes intensive treatment difficult. Several cases of donor-derived secondary leukemia in patients who received allogeneic hematopoietic stem cell transplantation (HSCT) have been reported [18,69,70,71]; thus, treatment decisions require the careful consideration of genetic background. Recent reports describe the development of acute lymphocytic leukemia and solid cancers in individuals with DDX41 variants [72,73], but the extent to which DDX41 variants are involved in such diseases remains controversial [64].
DDX41 has been shown to be essential for hematopoiesis, with homozygous Ddx41 knockout mice being embryonic lethal, although heterozygous mice show no remarkable abnormalities [74,75]. Several mechanisms have been proposed for the actions of DDX41 variants in the development of myeloid neoplasms. It has been reported that R-loop, a nucleic acid structure on the genome consisting of a DNA/RNA hybrid and single-strand DNA, aberrantly accumulates in MDS with RNA splicing abnormalities, regardless of the type of responsible gene [76,77,78,79], and that R-loop accumulation causes DNA replication stress, DNA damage, and abnormal mitosis. Recently, DDX41 has also been shown to be involved in R-loop regulation [80,81,82], and it is suggested that R-loop accumulation due to dysfunction or decreased expression of DDX41 is involved in impaired hematopoiesis and aberrant innate immune responses (Figure 1B). One of the major functions of DDX41 is RNA splicing [19]. However, considering that DDX41 variants develop de novo AML in addition to MDS, DDX41 is thought to play different roles from those of typical RNA splicing factors associated with MDS development. Indeed, while SRSF2, SF3B1, and U2AF1 are all involved in the recognition of pre-mRNA 3′ splice sites with U2 snRNP [83], DDX41 has been shown to be incorporated into the spliceosome at the C complex stage, a late complex of the activated spliceosome [82,84]. Regarding the relationship between DDX41 and R-loops, there are reports showing that DDX41 can unwind R-loops on its own [81,85], while it has also been suggested that impaired DDX41 function leads to reduced efficiency of RNA splicing, thus resulting in conditions that facilitate R-loop formation [82]. The accumulation of R-loop has been shown to give rise to an excessive innate immune reaction mediated through the cGAS-STING signaling pathway, consequently inducing increased hematopoietic stem/progenitor cells [80]. However, the mechanisms by which R-loops activate the cGAS-STING pathway remain inconclusive. Recently, it was reported that DNA/RNA hybrids derived from R-loops are transported to the cytoplasm and thus trigger an innate immune response [86]. The relevance of this observation to impaired hematopoiesis caused by DDX41 variants is of interest.
DDX41 is also reported to promote the processing of small nucleolar RNA (snoRNA) from introns [75]. Some snoRNA are coded within introns of ribosomal protein genes and mature after being processed from the introns [87,88]. snoRNAs are classified into boxC/D type and boxH/ACA types depending on their sequences; the former catalyzes 2′-O-methylation and the latter is responsible for catalyzing the pseudouridylation of uridine residues in ribosomal RNA, thereby promoting ribosomal biogenesis. Thus, loss of function (LOF) or expression of DDX41 impairs ribosomal biogenesis [66,89]. Although the involvement of DDX41 in ribosomal biogenesis has been reported by other research groups, the process involving DDX41 may be different from processes involving snoRNA processing.
Recently, myeloid neoplasms with germline DDX41 variants were shown to have a higher proportion of somatic CUX1 variants compared with those without a known germline background [61]. CUX1 is a transcription factor [90] that has also been shown to be directly involved in DNA damage repair by recruiting histone-modifying enzymes to damaged DNA regions [91]. Given that cells lacking sufficient CUX1 function can enter mitosis without completing DNA damage repair, the likelihood that the loss of DDX41 function or expression causes DNA replication stress is further increased. However, further studies are clearly needed to fully elucidate the mechanisms by which DDX41 variants lead to myeloid neoplasms.

2.2. Li-Fraumeni Syndrome (LFS)

TP53 is one of the most frequently mutated genes, especially in adult-onset cancers. Genome sequencing of various human cancer cells has revealed that 42% of cases carry TP53 variants [92]. The p53 protein is a transcription factor that can activate the expression of multiple target genes, plays an important role in the regulation of the cell cycle, apoptosis, and genomic stability, and is widely known as “the guardian of the genome”(Figure 2) [93,94]. The evidence accumulated to date suggests that p53 also regulates cell metabolism, ferroptosis, tumor microenvironment, and autophagy, which each contribute to tumor suppression [94]. Genomic instability caused by deletions and variants in TP53 may lead to accumulated gene mutations, causing gain of function (GOF) in the oncogene and LOF in the tumor suppressor gene [95]. p53 variants confer metabolic plasticity to cancer cells, promoting adaptation to metabolic stress and increasing the possibility of proliferation and metastasis [96].
The major type of TP53 variant is a missense variant producing a single amino acid substitution, with the DNA-binding domain (DBD) being the most mutated region [97]. Structural variants can reduce the thermostability of the protein, resulting in protein misfolding at physiological temperatures and a loss of its ability to bind DNA [98]. These variants not only bind wild-type p53 and cause dominant-negative (DN) effects, but may also be converted to oncogenic proteins via GOF, promoting various cellular responses such as carcinogenesis, cancer cell proliferation, invasion, metastasis, tumor microenvironment establishment, genomic instability, and metabolic reprogramming [99,100]. p53 is mutated and inactivated in most malignancies, making it a very attractive target for the development of new anti-cancer drugs [101]. Until recently, however, p53 was considered an undruggable target, and the progress made in p53-targeted therapeutics has been limited.
LFS is caused by a germline variant in the TP53 gene and is characterized by an increased risk of developing various solid tumors and hematologic malignancies at a young age [102,103]. LFS affects all ethnicities and has an estimated incidence of 1:5000 [27]. LFS is inherited in an autosomal dominant manner, although de novo inheritances occur in 7–20% of cases. Nearly 100% of individuals develop cancer by the age of 70, with the median age of first cancer at 20 to 30 years [26]. The tumor spectrum includes soft-tissue sarcomas, premenopausal breast cancer, central nervous system tumors, adrenocortical carcinomas, and pancreatic tumors, as well as MDS and lymphoid and myeloid malignancies. Germline TP53 variants are found in approximately 50% of pediatric patients with hypoploid acute lymphoblastic leukemia (ALL) and are associated with poor outcomes [104,105]. In the Le-Fraumeni lineage, leukemia is relatively uncommon, with only approximately 4% of children and adolescents presenting with hypodiploid ALL, treatment-related, or de novo MDS/AML [29].
As causal therapy is not available, the primary focus for improving the prognosis is early cancer detection. To this end, current cancer surveillance recommendations include a series of examinations including regular imaging beginning at birth [102]. As radiation exposure may lead to an increased (secondary) tumor risk, computed tomography and X-ray examinations should be avoided for as long as possible. Because annual whole-body magnetic resonance imaging has no radiation exposure and yet a high sensitivity for many tumors, it forms the basis of the recommended imaging [102].

2.3. AML with a Germline CEBPA Variant

The CEBPA gene is located on chromosome 19q13.1 and gene variants are a common genetic alteration in AML. Patients present with de novo AML [French American-British (FAB) classification; AML M1, M2, and M4 subtypes] and a group of differentiation abnormalities [106].
The single-exon gene CEBPA encodes CEBPa, which is the founder of the 6-CEBP family of transcription factors (TFs) [107]. All CEBP TFs contain a basic leucine zipper (bZIP) domain at the C terminus and form a subgroup within the leucine zipper family of TFs [108]. The CEBPa zipper domain is required for dimerization, and the adjacent basic region is responsible for DNA binding, thereby promoting the transcription of target genes [109]. The N terminus is unique to CEBPa, containing two transactivation domains that regulate transcription control and protein interaction [109]. CEBPa generates two isoforms from alternative initiation codons: the long isoform (p42) is 358 aa, and the short isoform (p30) is 239 aa and lacks a transactivation domain [110]. The p30 isoform maintains dimerization and DNA binding capacities, and, hence, can inhibit p42 activity. Both isoforms are coexpressed in a range of tissues, with p42 generally being more abundant [110]. Germline and somatic variants in CEBPA are clustered at the N terminus or within the C-terminal bZIP domain. These germline variants are generally frameshift or nonsense variants near the amino terminus of the encoded protein. Somatic variants in CEBPA often occur in the other allele, leading to a biallelic variant in CEBPA. This triggers the development of AML [31]. Commonly, the germline variant affects the N terminus, whereas the acquired variant arises in the C-terminal bZIP region [108]. The CEBPA variants that predispose to AML are generally considered to have a dominant-negative effect. The N-terminal truncating variants destroy p42, and the C-terminal variants abolish DNA binding or dimerization [108].
CEBPA-associated familial AML is defined as the presence of heterozygous germline CEBPA pathogenic variants in AML patients and/or in families with one or more AML patients. In contrast, sporadic CEBPA-associated AML is defined as AML in which the CEBPA pathogenic variant is identified in leukemic cells and not in non-leukemic cells [111]. AML with germline CEBPA variants generally occurs in autosomal-dominant inheritance without preceding abnormal blood cell counts or myelodysplasia [112]. Approximately 10% of CEBPA-associated AMLs have been shown to carry germline CEBPA variants [2]. In contrast to the incomplete penetrance observed in other HHMSs, germline CEBPA variants cause AML with almost complete penetrance (lifetime risk estimated to be >80%) [113]. Less than 20 families have been reported to have germline CEBPA variants [32]. In the majority of CEBPA-associated familial AML, the age of onset appears to be earlier than in sporadic CEBPA-associated AML [111]. Onset usually occurs in the 20th or 30th year of life, and many patients develop AML before 50 years of age; the median age of onset for AML is 24.5 years [34]. The prognosis of CEBPA-associated familial AML appears to be better than that of sporadic CEBPA-associated AML [114,115]. Patients with CEBPA-associated familial AML with a cured initial presentation are at high risk of developing additional independent leukemic episodes in addition to the risk of relapse from a pre-existing clone; the clinical observation that AML patients with CEBPA variants are more likely to develop a secondary leukemia despite their favorable prognosis is likely due to this pattern of progression [37]. Lifelong surveillance is recommended in patients with familial AML because of the high risk of late leukemia relapse [16]. It is important to avoid the use of allogeneic or consanguineous donors for HSCT without prior evaluation of the donor’s germline CEBPA pathogenic variant [116].

2.4. Myeloid Neoplasms with Other Germline Variants (ATM and CHEK2)

Deficiencies in the homologous recombination (HR) pathway can lead to defective DNA damage responses, and this can occur through inherited germline variants in HR pathway genes, such as checkpoint kinase 2 gene (CHEK2) and the ataxia telangiectasia mutated gene (ATM). The proper repair of DNA double-strand breaks (DSBs) is a core element of the maintenance of genomic stability, directed through three pathways active in most human cells: (1) homologous recombination (HR); (2) canonical non-homologous end joining (NHEJ); and (3) alternative NHEJ [117]. Canonical NHEJ is the simplest DNA repair mechanism. It involves directly adjoining DSBs through the binding of the Ku-80-Ku7p proteins to the fragmented DNA ends, followed by the recruitment of DNA-dependent protein kinases, which then activate ligase IV and co-factors which seal the DNA break. The alternative NHEJ mechanism involves the recruitment of PARP to the DNA ends, ending in the DNA DSB being sealed by Ligase I and III [118]. Although NHEJ effectively repairs DNA DSBs, it does not involve the usage of a complementary DNA template, and, as such, is error-prone, inducing chromosomal abnormalities and chromothripsis [118]. In contrast, HR is the most error-free of the DNA repair pathways, since it uses a complementary DNA template available during S-phase to correct the detected DNA lesion [119]. The HR pathway is engaged when the MRE11-RAD50-NBS1 protein complex is recruited to the fragmented DNA ends, which subsequently recruits ATM serine/threonine kinase. Activated ATM then phosphorylates the checkpoint kinase 2 (CHK2) protein, resulting in the downstream activation of a series of proteins, including CDC25C, p53, BRCA1/2, and cyclin-D kinases, which coordinate template-based DSB repair, cell-cycle arrest, and potentially apoptosis [117,118]. Another important feature of the DSB response is the induction of cell cycle checkpoint arrest, mediated during the S or G2 phase by ATR serine/threonine kinase and ATM, and it is essential for allowing the cell to re-enter mitosis after successful DSB repair [118].
Germline pathogenic and likely pathogenic gene variants that result in loss of function (LOF), such as ATM and CHEK2 variants, have been identified and characterized at several levels of the HR pathway. The broad consequence of a variant in one of these genes is a defective HR pathway, with consequent reliance on error-prone NHEJ mechanisms for DNA repair. The downstream result of using error-prone DNA repair pathways is an accumulation of somatic chromosomal abnormalities and DNA changes, particularly within rapidly dividing cells (e.g., epithelial, mammary, and hematopoietic), with an increased risk for the development of overt malignancy. Germline mutations in these genes have been well characterized as risk factors for breast [120], prostate [121,122], and pancreatic [123] cancers.
CHK2 protein is essential to the transmission of the DSB signal from ATM to downstream effectors CDC25C, p53, BRCA1/2, cyclin-D kinases, and others via phosphorylation. A variety of mutation types in CHEK2 have been identified, including splice site, missense, and frameshift, without a predisposition towards mutational hotspots [124]. Although the majority of patients carrying CHEK2 variants are in the heterozygous state, individuals with homozygous LOF CHEK2 do occur and have a Li-Fraumeni like phenotype. Heterozygous LOF CHEK2 variants are moderate penetrance risk factors for solid organ malignancies, including breast [125], prostate [126], renal cell carcinoma [127], papillary thyroid cancer [128], colorectal cancer [129], and pancreatic cancer [130]. LOF CHEK2 variants are increasingly recognized as risk factors for myeloid malignancies, including MPNs [131,132,133], MDS [134,135], and AML [136,137,138]. Germline CHEK2 variants have also been identified as risk factors for lymphoid malignancies [139].
Germline LOF mutations in the ATM gene have long been associated with early-onset myeloid malignancies, in addition to solid tumors such as breast and pancreatic cancers [140]. Loss of ATM function generates a greater risk of chromosomal translocations and other deleterious mutations associated with myeloid leukemia development [141]. Patients carrying LOF ATM variants in the homozygous or compound heterozygous states present with Ataxia Telangiectasia (A-T), an autosomal recessive disorder characterized by a 50- to 150-fold increased risk of cancer development, and also cerebellar degeneration, telangiectasia, immunodeficiency, and radiation sensitivity [142]. However, most individuals with germline deleterious ATM variants are heterozygous carriers with a 2- to 13-fold increased risk for early-onset cancer development but do not have other features of A-T [143]. The role for ATM mutations in myeloid malignancies remains in evolution and is less well characterized than for CHEK2. However, pathogenic ATM variants have been identified at diagnosis in several patients with de novo AML [62]. Intact ATM function has been well established as being critical for hematopoietic stem cell function [144], and ATM function and the associated signaling axis have been shown in vitro to modulate pathogenesis in AML [145,146]. In contrast to myeloid malignancies, germline ATM variants, either in the heterozygous state or in the context of true A-T with biallelic ATM variants, have been strongly associated with the development of lymphoid malignancies [147,148,149].

3. Myeloid Neoplasms with Preexisting Platelet Disorders

Most predisposition syndromes are associated with specific hematopoietic cell lineage abnormalities, and each exhibits a different tumor profile. For example, germline variants in RUNX1, ANKRD26, and ETV6 all predispose to thrombocytopenia and hematologic malignancies [150]. However, there are marked differences in cancer predisposition: the ANKRD26 variant predisposes to myeloid malignancies, ETV6 predominantly predisposes to B-cell ALL, and RUNX1 is associated with myeloid malignancies, and, to a lesser extent, predisposes to T-cell ALL [151]. Three different types of germ cell lineage predisposition are associated with highly variable penetrance in both myeloid and lymphoid systems. In both myeloid and lymphoid leukemias, the disease phenotype is likely influenced by both intrinsic and extrinsic cellular factors [150].

3.1. Myeloid and Lymphoid Neoplasms with a Germline RUNX1 Variant

RUNX1 encodes a heterodimeric transcription factor essential for hematopoiesis, megakaryopoiesis, and platelet function [152]. It functions as a transcriptional activator for some genes and a transcriptional repressor for others. Somatic variants in RUNX1 are among the most common variants in adults and children with ALL, AML, or MDS, including recurrent fusions in B-ALL (ETV6-RUNX1) and AML (RUNX1-RUNX1T1) [41]. RUNX1 was identified as a gene located at a truncation site on chromosome 21 in t (8;21), which is found in AML [153]. Somatic variants in the RUNX1 gene are one of the most frequently identified variants and have been identified in patients with various myeloid malignancies, including MDS, MPN, and AML [40]. In most cases, these RUNX1 variants are considered “subclonal variants” [154]. A high frequency of RUNX1 variants (30–50%) has been reported in treatment-related and radiation-related MDS and AML [155,156]. It is generally believed that RUNX1 variants lead to a loss of RUNX1 function [157]. In contrast, germline variants in the RUNX1 gene cause familial myeloid malignant platelet disorders (FPD/AML) with autosomal dominant inheritance, typically presenting with quantitative/qualitative platelet defects and a predisposition to myeloid malignancies like MDS and AML [158]. In this case, heterozygous inherited RUNX1 variants play a fundamental role in the etiology of FPD/AML [159]. However, these inherited RUNX1 variants are not sufficient to cause leukemia. It is thought that the accumulation of various variants, such as the CDC25C biallelic RUNX1 variant, and the TET2 variant, causes progression to preleukemic clones and eventually leads to the development of hematologic malignancies [38,160].
Germline variants in RUNX1 are among the most frequently detected variants in the pathogenesis of HHMS [38]; the RUNX1 gene encodes a DNA-binding subunit that contains a highly conserved runt-homology domain (RHD) for sequence-specific DNA binding [161]. Truncation lesions occur throughout the gene, but missense variants within the RHD are the most common. Others include nonsense, frameshifts, duplications, partial or total gene deletions, and gene rearrangements. Many RUNX1 variants cause haploinsufficiency [157]. RUNX1 variants cause defects in hematopoietic differentiation, resulting in decreased hematopoietic progenitor cell numbers and abnormal megakaryocyte differentiation. Tumorigenesis is most commonly caused by the somatic second hit of RUNX1. Typical clinical features of FPD/AML are gradual thrombocytopenia, aspirin-like qualitative platelet abnormalities, and a tendency to develop hematologic tumors [162].
Approximately 20–60% of FPD/AML families develop hematologic neoplasms during their lifetime [162]. More than 250 families have been reported to have germline RUNX1 variants. The latency period to transformation is relatively long, with the average age at diagnosis reported to be 33 years (maximum 76 years) [41]. Similar to what is observed in sporadic hematologic malignancies, additional acquired genetic events cooperate with the hereditary RUNX1 variant to progress the manifestation of the malignant phase. A comparative international cohort of germline RUNX1 variant carriers without and with hematological malignancies (HM) identified striking heterogeneity in rates of early-onset clonal hematopoiesis (CH), with a high prevalence of CH in RUNX1 carriers who did not have malignancies (carriers without HM). In RUNX1 carriers without HM with CH, TET2, PHF6, and BCOR were reported to be recurrently mutated in RUNX1-driven malignancies, suggesting that CH is a direct precursor to malignancy in RUNX1-driven HHMS [163].
Although most cases develop MDS or AML, other phenotypes have also been reported, including secondary leukemia, T-cell acute lymphoblastic leukemia (T-ALL), and non-Hodgkin lymphoma (NHL) [162]. Interestingly, the location of variants within the RUNX1 gene does not seem to affect disease phenotype among individuals, and phenotypic heterogeneity is often observed even within families with lesions of the same germ lineage [93].

3.2. Myeloid Neoplasms with a Germline ANKRD26 Variant

ANKRD26 is a gene located at 10p12.1 that regulates megakaryocyte development and thrombocytopenia [164]. RUNX1 and FLi1 co-regulate ANKRD26 by binding to the ANKRD26 promoter and repressing gene activity [165]. ANKRD26-related thrombocytopenia (ANKRD26 RT) is an autosomal dominant thrombocytopenia caused by a single nucleotide substitution in the ANKRD26 gene, characterized by quantitative and qualitative platelet disorders and an increased risk of MDS and AML [166]. ANKRD26 encodes a protein with an ankyrin repeat domain at its N-terminus and is thought to function in protein–protein interactions; while the function of the ANKRD26 protein is unknown, expression profiling has demonstrated its presence in megakaryocytes [166]. Germline variants in ANRK26 are usually point mutations located in the 5′ untranslated region (UTR) of the gene, although deletions and point mutations within the coding region have also been reported [167]. Variants in the 5′UTR affect the binding of repressive transcription factors such as RUNX1 and FLi1 to this regulatory region, abnormally increasing the expression of ANKRD26 and impairing platelet production [150]. The age of diagnosis generally ranges from early 20s to 70s. The incidence of myeloid malignancies is high in these patients, with an estimated 5% for AML, 2.2% for MDS, and 1.3% for chronic myeloid leukemia, with an estimated risk of these malignancies of 23, 12, and 21 times that of the general population, respectively [14].

3.3. Myeloid and Lymphoid Neoplasms with a Germline ETV6 Variant

Patients with thrombocytopenia 5, an autosomal dominant disorder of thrombocytopenia with bleeding tendency, usually present in childhood and have been found to have germline variants in ETV6 [168]. Clinical features include thrombocytopenia, abnormal platelet function, and increased bleeding tendency [49]. Leukemia is estimated to occur in about 30% of carriers, most commonly in ALL, but more than 30 translocation partners of ETV6 have been reported in AML, MDS, MPN, and T-cell lymphomas. ETV6 is one of the most commonly translocated genes in human AL and MDS [169]. ALL is more frequent, especially in B-ALL (0.8% of unselected childhood B-cell ALL). The ratio of lymphoid versus myeloid malignancies is roughly 2:1. Age ranges from 8 to 82 years and it seems to occur at a younger age than usual but is not yet defined [45,49,50].
ETV6 is located on chromosome 12p13.2 and encodes a transcriptional repressor important for hematopoiesis, megakaryopoiesis, and embryogenesis, and it is involved in angiogenesis, cell growth, and differentiation [170]. The gene encodes an N-terminal or C-terminal zinc finger, but the majority of variants are clustered within the DNA-binding ETS domain. Somatic rearrangements (most commonly with RUNX1), deletions, and sequence variants are observed in ALL. Second-hit variants (especially deletions) in ETV6 are common in ETV6-RUNX1 rearranged leukemias [171]. In addition, somatic rearrangements with RUNX1 are observed in a quarter of ALL patients [172]. Studies using umbilical cord blood from healthy newborns have shown that ETV6-RUNX1 translocations can occur in more than 1% of the healthy population [173].

4. Myeloid Neoplasms with Other Organ Dysfunction

4.1. Myeloid Neoplasms with a Germline SAMD9/SAMD9L Variant

SAMD9 and SAMD9L are a homologous gene pair at the head and tail of 7q21 and are interferon-inducible genes that are widely expressed in human tissues [55,174]. Both negatively regulate cell proliferation and function as tumor suppressors. Genetic variants in SAMD9/SAMD9L were initially shown to cause multisystem syndromes characterized by various neurological and/or endocrine abnormalities, as well as MDS with monosomy 7 and del7q [55,175]. Little is known about the biochemical activity of the SAMD9 and SAMD9L proteins and their domain structures, but they cluster in the latter half of the protein, in or near the putative P-loop [176]. The SAMD9 and SAMD9L proteins appear to be involved in endocytosis and cytokine signaling [177,178]; moreover, they have been reported to play a role in antiviral responses, similar to DDX41. Specifically, SAMD9 and SAMD9L are known to be host-restricted factors in poxvirus infection [179,180].
Germline variants in these genes are strongly associated with monogenic and familial pediatric MDS and potential full or partial deletions of adult chromosome 7(Figure 3) [53]. Germline variants in SAMD9 or SAMD9L are heterozygous gain-of-function missense variants, leading to proliferative arrest when expressed exogenously in the cell [174]. Carriers are at high risk for MDS and AML with cytopenia and monosomy 7/del7q. Many other patients who do not develop monosomy 7 acquire somatic variants in SAMD9 or SAMD9L resulting in the loss of function of the mutant protein [181]. The overexpression of SAMD9 or SAMD9L results in decreased proliferation and increased apoptosis, ultimately leading to the hypocellular phenotype being observed in patients. The effects on ribosome biology, DNA damage, and the resulting genomic instability are thought to promote the observed apoptotic phenotype [182,183] and ultimately lead to reduced bone marrow cellularity. Unrepaired DNA defects in hematopoietic cells cause significant long-term functional disruption and are a major driving force for the accumulation of further variants, thus promoting clonal expansion and malignant transformation [184,185,186].
Germline variants in SAMD9 cause a syndrome represented by the acronym MIRAGE; MIRAGE syndrome is an autosomal-dominant multisystem disorder characterized by six core features [187,188,189,190,191]. The features include bone marrow failure, progression to MDS and AML, infection, intrauterine dysplasia, adrenal hypoplasia, genital abnormalities, and enteropathy (chronic diarrhea with colonic dilatation). Germline variants in SAMD9L cause ataxia-pancytopenia syndrome, an autosomal dominant disorder with early onset gait and balance disturbances, nystagmus, mild pyramidal signs, and marked cerebellar atrophy [192,193,194,195]. Hematologic abnormalities include pancytopenia, bone marrow failure, and progression to MDS and AML. Germline variants in these two genes are found in 8–17% of pediatric MDS cases and more than 110 individuals have been reported to carry these germline variants [55]. It occurs mainly in childhood, but the average age of onset is not yet defined.

4.2. Myeloid Neoplasms with a Germline GATA2 Variant

Hematological malignancies affecting either the lymphoid or the myeloid lineages involve epigenetic mutations or dysregulation in the majority of cases. These epigenetic abnormalities can affect regulatory elements in the genome, and, particularly, enhancers. Recently, large regulatory elements known as super-enhancers (SE), initially identified for their critical roles in the cell-type specific expression regulation of genes controlling cell identity, have been shown to also be involved in tumorigenesis in many cancer types and hematological malignancies via the regulation of numerous oncogenes. Enhancer and SE hijacking refers to a mechanism by which an abnormally overexpressed TF binds to an inactive or poised enhancer already located near a given oncogene, recruiting other factors and chromatin remodelers. This binding allows the aberrant activation of the considered enhancer/SE, and, thereby, upregulates its associated oncogene. An example of such enhancer hijacking is provided by AML with the GATA2 SE translocated near the EVI1 promoter. A single enhancer contained within this GATA2 SE is composed of MYB binding sites, strongly required for EVI1 overexpression in AML cells. In addition, the mutation of this MYB binding site within this specific SE leads to myeloid differentiation, as well as cell death [196,197].
GATA2 is a zinc finger transcription factor that plays important roles in hematopoiesis, the homeostasis of hematopoietic stem cells (HSC), and lymphocyte development, specifically interacting with RUNX1 to control HSC survival [198]. GATA2 haploinsufficiency is caused by a missense variant or deletion in the GATA2 located on chromosome 3q21.3 [199]. Other causative variants have been detected throughout the gene, including nonsense, frameshift, splice site, and synonymous variants that cause splice abnormalities, as well as variants that target enhancers deep within introns [200]. GATA2 haploinsufficiency is an autosomal dominant inherited bone marrow failure and immunodeficiency syndrome predisposing to MDS and AML. The syndrome results from loss-of-function variants or deletions in the GATA2 gene [201]. Notably, GATA2 deficiency syndromes (G2DSs) show marked heterogeneity in inter- and intra-familial phenotypes, all within the spectrum of the single condition G2DS [13,202].
Phenotypes range from isolated chronic neutropenia to MDS/AML, bone marrow failure, severe immunodeficiency, and alveolar proteinosis. Patients may present with isolated neutropenia and bone marrow failure without syndromic features or family history [203]. Atypical mycobacterial infections, viral, and fungal infections are common, often overlapping with prolonged neutropenia, monocytopenia, B-cell deficiency, NK-cell deficiency, monocytopenia with Mycobacterium avium complex (MonoMAC) syndrome, or dendritic cell-monocyte-B-NK lymphocyte (DCML) deficiency [204,205]. Other symptoms include sensorineural hearing loss and lymphoedema (Emberger syndrome) [206,207].
Of particular note is that MDS/AML may present with one or more of these features, either years before the onset of MDS/AML or in isolation with MDS/AML. MDS with germline GATA2 variants is often associated with monosomy 7/del7q(-7) or trisomy 8, especially in children and young adults [205,208]. A study of 426 pediatric MDS cases identified germline GATA2 variants in 37% of patients with primary MDS with 7 and 16% of MDS cases with trisomy 8 [209]. In contrast, no germline GATA2 variants were found in treatment-related MDS. There have been over 480 individuals identified carrying a pathogenic or likely pathogenic germline GATA2 variant with symptoms of G2DS, with 240 of these confirmed to be familial and 24 de novo [57]. For those that develop myeloid malignancy (75% of all carriers with G2DS disease symptoms), the median age of onset is 17 years (range 0–78 years) and myelodysplastic syndrome is the first diagnosis in 75% of these cases with acute myeloid leukemia in a further 9% [57].

5. IBMFS

Inherited bone marrow failure syndrome (IBMFS) is an inherited disease associated with decreased bone marrow cell production [210,211,212]. It is associated with a specific clinical phenotype and variable risk of developing MDS or AML. Traditionally, the distinction has been made based on the presence or absence of classical physical manifestations [213] such as abnormal nails, reticulate pigmentation of the skin, and oral leukoplakia in congenital dyskeratosis. Fanconi anemia (FA) [214,215,216], Diamond-Blackfan anemia (DBA) [217,218,219], dyskeratosis congenita (DC) [220,221,222], telomere biology disorders (TBDs) [223], and Schwachman-Diamond syndrome (SDS) [224] are well-known predisposing factors for MDS/AML and exhibit characteristic physical symptoms and signs.
FA is an X-linked or autosomal recessive disorder characterized by genomic instability, hypersensitivity to DNA cross-linking agents, bone marrow failure, and predisposition to hematologic malignancies and solid tumors [210,211,212]. Hematologic abnormalities vary and include cytopenia, erythrocytosis, hypocellular bone marrow with mild dysplasia, and bone marrow failure with an increased risk of MDS or AML. The incidence of leukemia is even higher in the FANCD1/BRCA2 subtype of FA, with most cases occurring at less than 5 years of age [225]. This clinically and genetically diverse syndrome is caused by germline mutations in any of at least 23 FA genes (FANCA-FANCW) that function cooperatively in DNA repair. The risk of progression to MDS or AML is very high (cumulative incidence of AML at age 50 years is 10% and MDS at age 50 years is 40%) [226]. Unlike other MDSs that are cured by HSCT, these patients have higher post-transplant morbidity and a higher risk of solid tumors compared to non-transplant patients.
DBA usually presents in infancy with macrocytic anemia and reticulocytopenia. Bone marrow histology usually shows aplasia of erythrocytes in normocytic bone marrow. Major causes of morbidity and mortality are associated with side effects of treatment and a long-term risk of malignancy [217,218,219]. X-linked variants in GATA1, which encodes a transcription factor important for erythropoiesis, are also a cause of DBA [227]. Disease mechanisms include p53-mediated apoptosis induced by ribosomal stress, increased cell death due to excess free heme with delayed globin production, increased autophagy, and translational changes in selective erythroid-specific transcripts such as GATA1 [228].
DC/TBDs encompass genetically heterogeneous disorders associated with impaired telomere maintenance [220,221,222,223]. They are often associated with hematologic complications such as bone marrow failure, MDS, and AML. The cumulative incidence of MDS in DC/TBDs is estimated to be 2% by age 50 [229]. DC/TBD is associated with many non-hematologic complications, particularly pulmonary fibrosis, liver function abnormalities, and vascular abnormalities. Screening for TBD involves assessing the telomere length of lymphocytes, and further genetic testing for specific gene mutations is diagnostically useful because telomere shortening can also be seen in other diseases [230]. Telomeres shorten as the DNA replication cycle progresses. A critical shortening of telomere length leads to senescence and cell death [231].
SDS is characterized by pancreatic exocrine dysfunction and other physical findings. The most common nonhematologic abnormality is neurologic decompensation, which may be mild or severe, transient or persistent [224]. Other hematologic complications include bone marrow failure, MDS, and AML. In a French cohort of 102 SDS patients, the cumulative incidence of MDS/AML was 18.8% at age 20 and 36.1% at age 30 [232]. SDS is most often caused by an autosomal recessive mutation in the eponymous SBDS gene, resulting in low levels of SDS protein. SDS is involved in the binding of the large and small ribosomal subunits and functions as an elongation factor-like cofactor that removes the anti-binding factor eukaryotic initiation factor 6 (eIF6) from the large subunit [233]. SDS is also involved in the stabilization of mitotic spindles. The spectrum of SBDS variants, including missense, splice site, nonsense, frameshift, and partial or total gene deletions, has been confirmed. AML has been reported in patients with variants in the autosomal recessive gene in DnaJ Heat Shock Protein Family Member C21 genes (DNAJC21) and in those with various clinical features of SDS [234].

6. Infant Leukemia with a Germline Predisposition

Some infant leukemias with a germline predisposition have been reported and elucidated, although the section on them is not added in the 2016 revision of the WHO classification. Pediatric cancers typically harbor relatively few somatic mutations and frequently demonstrate developmentally immature phenotypes, suggesting a contribution of germline variation that might result in aberrant tissue development [235]. MLL rearrangements are observed in approximately 50–80% of infant ALL cases and 34–50% of infant AML cases [236]. There is evidence from multiple in vitro systems that the presence of a MLL rearrangement is insufficient by itself to drive leukemogenesis [237,238,239], suggesting that additional factors are required in the presence (and absence) of MLL rearrangements to drive leukemogenesis.
KMT2 protein is an epigenetic modifier, and each histone modification is associated with regulatory elements and mediates specific functions, enabling complex control over gene transcription [240]. KMT2C and KMT2D play an essential role in mediating monomethylation at histone 3 and lysine 4, primarily at enhancers [241]. Germline or somatic variations in a family of KMT2 lysine methyltransferases have been associated with a variety of congenital disorders and cancers. In mammals, somatic mutations of KMT2C and KMT2D are associated with various malignancies [242], with clear evidence for tumor suppressor roles [243,244]. Notably, KMT2A-fusions are prevalent in 70% of infant leukemias but fail to phenocopy short latency leukemogenesis in mammalian models, suggesting additional factors are necessary for transformation [245]. Heterozygous germline missense variants in KMT2C are more common in infant leukemia compared to healthy controls [246]. The loss of KMT2C in mice leads to aberrant myelopoiesis, causing myeloid infiltration into lymphoid organs; however, the loss of KMT2C alone is insufficient to drive leukemia [247]. Somatic cell drivers such as KMT2A fusions added to germline KMT2C mutations may more readily transform hematopoietic progenitor cells.

7. Conclusions and Perspectives

As discussed above, the genetic and phenotypic background of HHMS has been rapidly elucidated over the past decade, and the disease is now diverse. Most HHMS-related genes have clearly defined functions that contribute to hematopoietic regulation. However, the precise nature of this association requires further investigation. Advances in HHMS practice have been made possible by the introduction of next-generation sequencing (NGS) technology in germline and somatic gene testing. These tests now often have overlapping gene lists and have gained international recognition, especially for the diagnosis and management of myeloid malignancy. The association between germline genes predisposing to solid tumors and hematologic tumors is also becoming clearer. For example, variants in breast cancer gene type 1/2 (BRCA1/2), partner and localizer of BRCA2 (PALB2), and TP53 occur in primary or treatment-related hematological malignancies, including AML, ALL, and MDS, narrowing the apparent distinction between solid tumors and hematologic tumor predisposition [248,249,250]. Future development of a hematologic tumor testing panel that is also useful in detecting refractory cytopenia and the risk of relapse refractoriness after leukemia-directed therapy is warranted. Extensive sequencing technologies, such as whole exome sequencing (WES), allow for the investigation of new candidate genetic abnormalities, including germline gene variants, at once, and are expected to be utilized more than targeted NGS panels in the future [4,251]. There is a growing need for the expert consultation and clinical surveillance of patients with a germline predisposition to hematologic malignancies [252]. Troublingly, prognosis and disease progression are slow. Therefore, consultation and treatment strategies must be tailored to the individual patient. Low-penetrance variants along with the late onset of the disease in some cases may be responsible for the delay in hereditary susceptibility recognition and have led some experts to propose universal germline testing strategies [248,253]. Bone marrow stromal cells show the advantage of being a readily available material from routine bone marrow aspirations, which can be isolated by culture [254]. However, confirmation in nonhematopoietic tissue or in other family members is necessary to avoid the misinterpretation of variants involved in CH, somatic copy number variants, or somatic loss heterozygosity [255]. For this purpose, skin fibroblasts are considered the gold standard, despite the requirement of a skin biopsy and long-lasting cultures. Patients and family members with suspected HHMS should be advised of the indications for genetic testing, the limitations of genetic testing, and genetic counseling. This is because curative therapy influences the outcome of allogeneic HSCT, regardless of the phenotypic spectrum or clinical presentation of HHMS [256]. The outcome in these patients is often poor, making them candidates for allogeneic HSCT. Compatible blood stem cell donors should be carefully considered, and donors with known germline variants or unknown retention status should be avoided. There are reports of cases of leukemia after allogeneic transplantation from blood donors [69]. DDX41, CEBPA, GATA2, and others have been reported to be present in 1~2% of allogeneic post-transplant relapses [257] with a median time of recurrence of 5.2 years [258]. There are also reports of onset 10 years after transplantation [70]. Various guidelines for genetic testing for HHMS are currently being proposed by organizations such as the National Comprehensive Cancer Network [259] and the American Society of Clinical Oncology [260]. However, rapid advances in the elucidation of the biology of hematologic tumors and in the clinical care of patients with these diseases necessitate the development of more detailed clinical guidelines. Providing clear eligibility criteria for HHMS testing, including the full spectrum of HHMS-related mutations, would improve the diagnosis and care of patients with these syndromes. Currently, no specific treatment for HHMS exists, and patients are not adequately treated. The lifelong surveillance of patients and their families is recommended to monitor for treatment-related toxicity, disease recurrence, and the development of new symptoms or signs in unaffected individuals. There is an international need to develop a comprehensive foundation for determining evidence-based management, family counseling, the treatment of symptomatic individuals, and preemptive interventions.

Author Contributions

H.A., H.M., S.C. and Y.M. were responsible for the preparation and writing of the manuscript; Y.U., S.M. and N.A. gave academic advice related to the theme of the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This paper was supported by a National Cancer Research and Development expenses grant (2021-A-11), funded by the National Cancer Center, Japan.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Corces-Zimmerman, M.R.; Majeti, R. Pre-leukemic evolution of hematopoietic stem cells: The importance of early mutations in leukemogenesis. Leukemia 2014, 28, 2276–2282. [Google Scholar] [CrossRef] [PubMed]
  2. Godley, L.A. Germline mutations in MDS/AML predisposition disorders. Curr. Opin. Hematol. 2021, 28, 86–93. [Google Scholar] [CrossRef] [PubMed]
  3. Fenwarth, L.; Caulier, A.; Lachaier, E.; Goursaud, L.; Marceau-Renaut, A.; Fournier, E.; Lebon, D.; Boyer, T.; Berthon, C.; Marolleau, J.P.; et al. Hereditary Predisposition to Acute Myeloid Leukemia in Older Adults. Hemasphere 2021, 5, e552. [Google Scholar] [CrossRef] [PubMed]
  4. Guijarro, F.; Lopez-Guerra, M.; Morata, J.; Bataller, A.; Paz, S.; Cornet-Masana, J.M.; Banus-Mulet, A.; Cuesta-Casanovas, L.; Carbo, J.M.; Castano-Diez, S.; et al. Germ line variants in patients with acute myeloid leukemia without a suspicion of hereditary hematologic malignancy syndrome. Blood Adv. 2023, 7, 5799–5811. [Google Scholar] [CrossRef] [PubMed]
  5. Churpek, J.E. Familial myelodysplastic syndrome/acute myeloid leukemia. Best Pract. Res. Clin. Haematol. 2017, 30, 287–289. [Google Scholar] [CrossRef] [PubMed]
  6. Hamidi, A.; Roloff, G.W.; Shaw, R.; Acevedo, M.; Smith, S.; Drazer, M.W. Clinical guideline variability in the diagnosis of hereditary hematopoietic malignancy syndromes. Leuk. Lymphoma 2023, 64, 1562–1565. [Google Scholar] [CrossRef]
  7. Stieglitz, E.; Loh, M.L. Genetic predispositions to childhood leukemia. Ther. Adv. Hematol. 2013, 4, 270–290. [Google Scholar] [CrossRef] [PubMed]
  8. Babushok, D.V.; Bessler, M. Genetic predisposition syndromes: When should they be considered in the work-up of MDS? Best Pract. Res. Clin. Haematol. 2015, 28, 55–68. [Google Scholar] [CrossRef]
  9. Kotmayer, L.; Kallay, K.; Bodor, C. Hereditary haematological malignancies. Magy. Onkol. 2020, 64, 43–55. [Google Scholar]
  10. Furutani, E.; Shimamura, A. Germline Genetic Predisposition to Hematologic Malignancy. J. Clin. Oncol. 2017, 35, 1018–1028. [Google Scholar] [CrossRef]
  11. Zahid, M.F.; Malik, U.A.; Sohail, M.; Hassan, I.N.; Ali, S.; Shaukat, M.H.S. Cytogenetic Abnormalities in Myelodysplastic Syndromes: An Overview. Int. J. Hematol. Oncol. Stem Cell Res. 2017, 11, 231–239. [Google Scholar] [PubMed]
  12. Yoshida, M.; Tanase-Nakao, K.; Shima, H.; Shirai, R.; Yoshida, K.; Osumi, T.; Deguchi, T.; Mori, M.; Arakawa, Y.; Takagi, M.; et al. Prevalence of germline GATA2 and SAMD9/9L variants in paediatric haematological disorders with monosomy 7. Br. J. Haematol. 2020, 191, 835–843. [Google Scholar] [CrossRef] [PubMed]
  13. Sahoo, S.S.; Kozyra, E.J.; Wlodarski, M.W. Germline predisposition in myeloid neoplasms: Unique genetic and clinical features of GATA2 deficiency and SAMD9/SAMD9L syndromes. Best Pract. Res. Clin. Haematol. 2020, 33, 101197. [Google Scholar] [CrossRef] [PubMed]
  14. Rafei, H.; DiNardo, C.D. Hereditary myeloid malignancies. Best Pract. Res. Clin. Haematol. 2019, 32, 163–176. [Google Scholar] [CrossRef] [PubMed]
  15. Zhang, M.Y.; Churpek, J.E.; Keel, S.B.; Walsh, T.; Lee, M.K.; Loeb, K.R.; Gulsuner, S.; Pritchard, C.C.; Sanchez-Bonilla, M.; Delrow, J.J.; et al. Germline ETV6 mutations in familial thrombocytopenia and hematologic malignancy. Nat. Genet. 2015, 47, 180–185. [Google Scholar] [CrossRef] [PubMed]
  16. Tawana, K.; Brown, A.L.; Churpek, J.E. Integrating germline variant assessment into routine clinical practice for myelodysplastic syndrome and acute myeloid leukaemia: Current strategies and challenges. Br. J. Haematol. 2022, 196, 1293–1310. [Google Scholar] [CrossRef] [PubMed]
  17. Arber, D.A.; Orazi, A.; Hasserjian, R.; Thiele, J.; Borowitz, M.J.; Le Beau, M.M.; Bloomfield, C.D.; Cazzola, M.; Vardiman, J.W. The 2016 revision to the World Health Organization classification of myeloid neoplasms and acute leukemia. Blood 2016, 127, 2391–2405. [Google Scholar] [CrossRef] [PubMed]
  18. Berger, G.; van den Berg, E.; Sikkema-Raddatz, B.; Abbott, K.M.; Sinke, R.J.; Bungener, L.B.; Mulder, A.B.; Vellenga, E. Re-emergence of acute myeloid leukemia in donor cells following allogeneic transplantation in a family with a germline DDX41 mutation. Leukemia 2017, 31, 520–522. [Google Scholar] [CrossRef]
  19. Polprasert, C.; Schulze, I.; Sekeres, M.A.; Makishima, H.; Przychodzen, B.; Hosono, N.; Singh, J.; Padgett, R.A.; Gu, X.; Phillips, J.G.; et al. Inherited and Somatic Defects in DDX41 in Myeloid Neoplasms. Cancer Cell 2015, 27, 658–670. [Google Scholar] [CrossRef]
  20. Lewinsohn, M.; Brown, A.L.; Weinel, L.M.; Phung, C.; Rafidi, G.; Lee, M.K.; Schreiber, A.W.; Feng, J.; Babic, M.; Chong, C.E.; et al. Novel germ line DDX41 mutations define families with a lower age of MDS/AML onset and lymphoid malignancies. Blood 2016, 127, 1017–1023. [Google Scholar] [CrossRef]
  21. Cheah, J.J.C.; Hahn, C.N.; Hiwase, D.K.; Scott, H.S.; Brown, A.L. Myeloid neoplasms with germline DDX41 mutation. Int. J. Hematol. 2017, 106, 163–174. [Google Scholar] [CrossRef] [PubMed]
  22. Makishima, H.; Bowman, T.V.; Godley, L.A. DDX41-associated susceptibility to myeloid neoplasms. Blood 2023, 141, 1544–1552. [Google Scholar] [CrossRef] [PubMed]
  23. Sébert, M.; Passet, M.; Raimbault, A.; Rahmé, R.; Raffoux, E.; Sicre de Fontbrune, F.; Cerrano, M.; Quentin, S.; Vasquez, N.; Da Costa, M.; et al. Germline DDX41 mutations define a significant entity within adult MDS/AML patients. Blood 2019, 134, 1441–1444. [Google Scholar] [CrossRef] [PubMed]
  24. Lalloo, F.; Varley, J.; Ellis, D.; Moran, A.; O’Dair, L.; Pharoah, P.; Evans, D.G.; The Early Onset Breast Cancer Study Group. Prediction of pathogenic mutations in patients with early-onset breast cancer by family history. Lancet 2003, 361, 1101–1102. [Google Scholar] [CrossRef] [PubMed]
  25. Bougeard, G.; Renaux-Petel, M.; Flaman, J.M.; Charbonnier, C.; Fermey, P.; Belotti, M.; Gauthier-Villars, M.; Stoppa-Lyonnet, D.; Consolino, E.; Brugieres, L.; et al. Revisiting Li-Fraumeni Syndrome from TP53 Mutation Carriers. J. Clin. Oncol. 2015, 33, 2345–2352. [Google Scholar] [CrossRef] [PubMed]
  26. Shin, S.J.; Dodd-Eaton, E.B.; Peng, G.; Bojadzieva, J.; Chen, J.; Amos, C.I.; Frone, M.N.; Khincha, P.P.; Mai, P.L.; Savage, S.A.; et al. Penetrance of Different Cancer Types in Families with Li-Fraumeni Syndrome: A Validation Study Using Multicenter Cohorts. Cancer Res. 2020, 80, 354–360. [Google Scholar] [CrossRef] [PubMed]
  27. Frebourg, T.; Bajalica Lagercrantz, S.; Oliveira, C.; Magenheim, R.; Evans, D.G.; European Reference Network GENTURIS. Guidelines for the Li-Fraumeni and heritable TP53-related cancer syndromes. Eur. J. Hum. Genet. 2020, 28, 1379–1386. [Google Scholar] [CrossRef] [PubMed]
  28. Winter, G.; Kirschner-Schwabe, R.; Groeneveld-Krentz, S.; Escherich, G.; Moricke, A.; von Stackelberg, A.; Stanulla, M.; Bailey, S.; Richter, L.; Steinemann, D.; et al. Clinical and genetic characteristics of children with acute lymphoblastic leukemia and Li-Fraumeni syndrome. Leukemia 2021, 35, 1475–1479. [Google Scholar] [CrossRef]
  29. Swaminathan, M.; Bannon, S.A.; Routbort, M.; Naqvi, K.; Kadia, T.M.; Takahashi, K.; Alvarado, Y.; Ravandi-Kashani, F.; Patel, K.P.; Champlin, R.; et al. Hematologic malignancies and Li-Fraumeni syndrome. Cold Spring Harb. Mol. Case Stud. 2019, 5, a003210. [Google Scholar] [CrossRef]
  30. Smith, M.L.; Cavenagh, J.D.; Lister, T.A.; Fitzgibbon, J. Mutation of CEBPA in familial acute myeloid leukemia. N. Engl. J. Med. 2004, 351, 2403–2407. [Google Scholar] [CrossRef]
  31. Pathak, A.; Seipel, K.; Pemov, A.; Dewan, R.; Brown, C.; Ravichandran, S.; Luke, B.T.; Malasky, M.; Suman, S.; Yeager, M.; et al. Whole exome sequencing reveals a C-terminal germline variant in CEBPA-associated acute myeloid leukemia: 45-year follow up of a large family. Haematologica 2016, 101, 846–852. [Google Scholar] [CrossRef] [PubMed]
  32. Mendoza, H.; Chen, P.H.; Pine, A.B.; Siddon, A.J.; Bale, A.E.; Gowda, L.; Killie, A.; Richards, J.; Varin-Tremblay, C.; Kloss, R.; et al. A case of acute myeloid leukemia with unusual germline CEBPA mutation: Lessons learned about mutation detection, location, and penetrance. Leuk. Lymphoma 2021, 62, 1251–1254. [Google Scholar] [CrossRef] [PubMed]
  33. Tawana, K.; Rio-Machin, A.; Preudhomme, C.; Fitzgibbon, J. Familial CEBPA-mutated acute myeloid leukemia. Semin. Hematol. 2017, 54, 87–93. [Google Scholar] [CrossRef] [PubMed]
  34. Harrigan, A.M.; Trottier, A.M. Hereditary acute myeloid leukemia associated with C-terminal CEBPA germline variants. Fam. Cancer 2023, 22, 331–339. [Google Scholar] [CrossRef] [PubMed]
  35. Jongmans, M.C.; Kuiper, R.P.; Carmichael, C.L.; Wilkins, E.J.; Dors, N.; Carmagnac, A.; Schouten-Van Meeteren, A.Y.; Li, X.; Stankovic, M.; Kamping, E.; et al. Novel RUNX1 mutations in familial platelet disorder with enhanced risk for acute myeloid leukemia: Clues for improved identification of the FPD/AML syndrome. Leukemia 2010, 24, 242–246. [Google Scholar] [CrossRef] [PubMed]
  36. Preudhomme, C.; Renneville, A.; Bourdon, V.; Philippe, N.; Roche-Lestienne, C.; Boissel, N.; Dhedin, N.; Andre, J.M.; Cornillet-Lefebvre, P.; Baruchel, A.; et al. High frequency of RUNX1 biallelic alteration in acute myeloid leukemia secondary to familial platelet disorder. Blood 2009, 113, 5583–5587. [Google Scholar] [CrossRef] [PubMed]
  37. Brown, A.L.; Hahn, C.N.; Scott, H.S. Secondary leukemia in patients with germline transcription factor mutations (RUNX1, GATA2, CEBPA). Blood 2020, 136, 24–35. [Google Scholar] [CrossRef] [PubMed]
  38. Brown, A.L.; Arts, P.; Carmichael, C.L.; Babic, M.; Dobbins, J.; Chong, C.E.; Schreiber, A.W.; Feng, J.; Phillips, K.; Wang, P.P.S.; et al. RUNX1-mutated families show phenotype heterogeneity and a somatic mutation profile unique to germline predisposed AML. Blood Adv. 2020, 4, 1131–1144. [Google Scholar] [CrossRef]
  39. Luo, X.; Feurstein, S.; Mohan, S.; Porter, C.C.; Jackson, S.A.; Keel, S.; Chicka, M.; Brown, A.L.; Kesserwan, C.; Agarwal, A.; et al. ClinGen Myeloid Malignancy Variant Curation Expert Panel recommendations for germline RUNX1 variants. Blood Adv. 2019, 3, 2962–2979. [Google Scholar] [CrossRef]
  40. Homan, C.C.; King-Smith, S.L.; Lawrence, D.M.; Arts, P.; Feng, J.; Andrews, J.; Armstrong, M.; Ha, T.; Dobbins, J.; Drazer, M.W.; et al. The RUNX1 database (RUNX1db): Establishment of an expert curated RUNX1 registry and genomics database as a public resource for familial platelet disorder with myeloid malignancy. Haematologica 2021, 106, 3004–3007. [Google Scholar] [CrossRef]
  41. Forster, A.; Decker, M.; Schlegelberger, B.; Ripperger, T. Beyond Pathogenic RUNX1 Germline Variants: The Spectrum of Somatic Alterations in RUNX1-Familial Platelet Disorder with Predisposition to Hematologic Malignancies. Cancers 2022, 14, 3431. [Google Scholar] [CrossRef] [PubMed]
  42. Noris, P.; Perrotta, S.; Seri, M.; Pecci, A.; Gnan, C.; Loffredo, G.; Pujol-Moix, N.; Zecca, M.; Scognamiglio, F.; De Rocco, D.; et al. Mutations in ANKRD26 are responsible for a frequent form of inherited thrombocytopenia: Analysis of 78 patients from 21 families. Blood 2011, 117, 6673–6680. [Google Scholar] [CrossRef] [PubMed]
  43. Pippucci, T.; Savoia, A.; Perrotta, S.; Pujol-Moix, N.; Noris, P.; Castegnaro, G.; Pecci, A.; Gnan, C.; Punzo, F.; Marconi, C.; et al. Mutations in the 5′ UTR of ANKRD26, the ankirin repeat domain 26 gene, cause an autosomal-dominant form of inherited thrombocytopenia, THC2. Am. J. Hum. Genet. 2011, 88, 115–120. [Google Scholar] [CrossRef] [PubMed]
  44. Boutroux, H.; Petit, A.; Auvrignon, A.; Lapillonne, H.; Ballerini, P.; Favier, R.; Leverger, G. Childhood diagnosis of genetic thrombocytopenia with mutation in the ankyrine repeat domain 26 gene. Eur. J. Pediatr. 2015, 174, 1399–1403. [Google Scholar] [CrossRef]
  45. Feurstein, S.; Godley, L.A. Germline ETV6 mutations and predisposition to hematological malignancies. Int. J. Hematol. 2017, 106, 189–195. [Google Scholar] [CrossRef]
  46. Bejar, R.; Stevenson, K.; Abdel-Wahab, O.; Galili, N.; Nilsson, B.; Garcia-Manero, G.; Kantarjian, H.; Raza, A.; Levine, R.L.; Neuberg, D.; et al. Clinical effect of point mutations in myelodysplastic syndromes. N. Engl. J. Med. 2011, 364, 2496–2506. [Google Scholar] [CrossRef]
  47. 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]
  48. Van Vlierberghe, P.; Ambesi-Impiombato, A.; Perez-Garcia, A.; Haydu, J.E.; Rigo, I.; Hadler, M.; Tosello, V.; Della Gatta, G.; Paietta, E.; Racevskis, J.; et al. ETV6 mutations in early immature human T cell leukemias. J. Exp. Med. 2011, 208, 2571–2579. [Google Scholar] [CrossRef]
  49. Noetzli, L.; Lo, R.W.; Lee-Sherick, A.B.; Callaghan, M.; Noris, P.; Savoia, A.; Rajpurkar, M.; Jones, K.; Gowan, K.; Balduini, C.; et al. Germline mutations in ETV6 are associated with thrombocytopenia, red cell macrocytosis and predisposition to lymphoblastic leukemia. Nat. Genet. 2015, 47, 535–538. [Google Scholar] [CrossRef]
  50. Moriyama, T.; Metzger, M.L.; Wu, G.; Nishii, R.; Qian, M.; Devidas, M.; Yang, W.; Cheng, C.; Cao, X.; Quinn, E.; et al. Germline genetic variation in ETV6 and risk of childhood acute lymphoblastic leukaemia: A systematic genetic study. Lancet Oncol. 2015, 16, 1659–1666. [Google Scholar] [CrossRef]
  51. Narumi, S.; Amano, N.; Ishii, T.; Katsumata, N.; Muroya, K.; Adachi, M.; Toyoshima, K.; Tanaka, Y.; Fukuzawa, R.; Miyako, K.; et al. SAMD9 mutations cause a novel multisystem disorder, MIRAGE syndrome, and are associated with loss of chromosome 7. Nat. Genet. 2016, 48, 792–797. [Google Scholar] [CrossRef] [PubMed]
  52. Chen, D.H.; Below, J.E.; Shimamura, A.; Keel, S.B.; Matsushita, M.; Wolff, J.; Sul, Y.; Bonkowski, E.; Castella, M.; Taniguchi, T.; et al. Ataxia-Pancytopenia Syndrome Is Caused by Missense Mutations in SAMD9L. Am. J. Hum. Genet. 2016, 98, 1146–1158. [Google Scholar] [CrossRef]
  53. Davidsson, J.; Puschmann, A.; Tedgard, U.; Bryder, D.; Nilsson, L.; Cammenga, J. SAMD9 and SAMD9L in inherited predisposition to ataxia, pancytopenia, and myeloid malignancies. Leukemia 2018, 32, 1106–1115. [Google Scholar] [CrossRef] [PubMed]
  54. Buonocore, F.; Kuhnen, P.; Suntharalingham, J.P.; Del Valle, I.; Digweed, M.; Stachelscheid, H.; Khajavi, N.; Didi, M.; Brady, A.F.; Blankenstein, O.; et al. Somatic mutations and progressive monosomy modify SAMD9-related phenotypes in humans. J. Clin. Investig. 2017, 127, 1700–1713. [Google Scholar] [CrossRef] [PubMed]
  55. Wong, J.C.; Bryant, V.; Lamprecht, T.; Ma, J.; Walsh, M.; Schwartz, J.; Del Pilar Alzamora, M.; Mullighan, C.G.; Loh, M.L.; Ribeiro, R.; et al. Germline SAMD9 and SAMD9L mutations are associated with extensive genetic evolution and diverse hematologic outcomes. JCI Insight 2018, 3, e121086. [Google Scholar] [CrossRef]
  56. Hahn, C.N.; Chong, C.E.; Carmichael, C.L.; Wilkins, E.J.; Brautigan, P.J.; Li, X.C.; Babic, M.; Lin, M.; Carmagnac, A.; Lee, Y.K.; et al. Heritable GATA2 mutations associated with familial myelodysplastic syndrome and acute myeloid leukemia. Nat. Genet. 2011, 43, 1012–1017. [Google Scholar] [CrossRef]
  57. Homan, C.C.; Venugopal, P.; Arts, P.; Shahrin, N.H.; Feurstein, S.; Rawlings, L.; Lawrence, D.M.; Andrews, J.; King-Smith, S.L.; Harvey, N.L.; et al. GATA2 deficiency syndrome: A decade of discovery. Hum. Mutat. 2021, 42, 1399–1421. [Google Scholar] [CrossRef]
  58. Al Seraihi, A.F.; Rio-Machin, A.; Tawana, K.; Bodor, C.; Wang, J.; Nagano, A.; Heward, J.A.; Iqbal, S.; Best, S.; Lea, N.; et al. GATA2 monoallelic expression underlies reduced penetrance in inherited GATA2-mutated MDS/AML. Leukemia 2018, 32, 2502–2507. [Google Scholar] [CrossRef]
  59. Bohnsack, K.E.; Yi, S.; Venus, S.; Jankowsky, E.; Bohnsack, M.T. Cellular functions of eukaryotic RNA helicases and their links to human diseases. Nat. Rev. Mol. Cell. Biol. 2023, 24, 749–769. [Google Scholar] [CrossRef]
  60. Fairman-Williams, M.E.; Guenther, U.P.; Jankowsky, E. SF1 and SF2 helicases: Family matters. Curr. Opin. Struct. Biol. 2010, 20, 313–324. [Google Scholar] [CrossRef]
  61. Makishima, H.; Saiki, R.; Nannya, Y.; Korotev, S.; Gurnari, C.; Takeda, J.; Momozawa, Y.; Best, S.; Krishnamurthy, P.; Yoshizato, T.; et al. Germ line DDX41 mutations define a unique subtype of myeloid neoplasms. Blood 2023, 141, 534–549. [Google Scholar] [CrossRef] [PubMed]
  62. Yang, F.; Long, N.; Anekpuritanang, T.; Bottomly, D.; Savage, J.C.; Lee, T.; Solis-Ruiz, J.; Borate, U.; Wilmot, B.; Tognon, C.; et al. Identification and prioritization of myeloid malignancy germline variants in a large cohort of adult patients with AML. Blood 2022, 139, 1208–1221. [Google Scholar] [CrossRef] [PubMed]
  63. Li, P.; Brown, S.; Williams, M.; White, T.; Xie, W.; Cui, W.; Peker, D.; Lei, L.; Kunder, C.A.; Wang, H.Y.; et al. The genetic landscape of germline DDX41 variants predisposing to myeloid neoplasms. Blood 2022, 140, 716–755. [Google Scholar] [CrossRef] [PubMed]
  64. Cheloor Kovilakam, S.; Gu, M.; Dunn, W.G.; Marando, L.; Barcena, C.; Nik-Zainal, S.; Mohorianu, I.; Kar, S.; Fabre, M.A.; Quiros, P.M.; et al. Prevalence and significance of DDX41 gene variants in the general population. Blood 2023, 142, 1185–1192. [Google Scholar] [CrossRef]
  65. Molteni, E.; Bono, E.; Galli, A.; Elena, C.; Ferrari, J.; Fiorelli, N.; Pozzi, S.; Ferretti, V.V.; Sarchi, M.; Rizzo, E.; et al. Prevalence and clinical expression of germline predisposition to myeloid neoplasms in adults with marrow hypocellularity. Blood 2023, 142, 643–657. [Google Scholar] [CrossRef] [PubMed]
  66. Kadono, M.; Kanai, A.; Nagamachi, A.; Shinriki, S.; Kawata, J.; Iwato, K.; Kyo, T.; Oshima, K.; Yokoyama, A.; Kawamura, T.; et al. Biological implications of somatic DDX41 p.R525H mutation in acute myeloid leukemia. Exp. Hematol. 2016, 44, 745–754.e4. [Google Scholar] [CrossRef]
  67. Tierens, A.; Kagotho, E.; Shinriki, S.; Seto, A.; Smith, A.C.; Care, M.; Maze, D.; Sibai, H.; Yee, K.W.; Schuh, A.C.; et al. Biallelic disruption of DDX41 activity is associated with distinct genomic and immunophenotypic hallmarks in acute leukemia. Front. Oncol. 2023, 13, 1153082. [Google Scholar] [CrossRef]
  68. Badar, T.; Nanaa, A.; Foran, J.M.; Viswanatha, D.; Al-Kali, A.; Lasho, T.; Finke, C.; Alkhateeb, H.B.; He, R.; Gangat, N.; et al. Clinical and molecular correlates of somatic and germline DDX41 variants in patients and families with myeloid neoplasms. Haematologica 2023, 108, 3033–3043. [Google Scholar] [CrossRef]
  69. Kobayashi, S.; Kobayashi, A.; Osawa, Y.; Nagao, S.; Takano, K.; Okada, Y.; Tachi, N.; Teramoto, M.; Kawamura, T.; Horiuchi, T.; et al. Donor cell leukemia arising from preleukemic clones with a novel germline DDX41 mutation after allogenic hematopoietic stem cell transplantation. Leukemia 2017, 31, 1020–1022. [Google Scholar] [CrossRef]
  70. Rolles, B.; Meyer, R.; Begemann, M.; Elbracht, M.; Jost, E.; Stelljes, M.; Kurth, I.; Brummendorf, T.H.; Silling, G. DDX41 germline variants causing donor cell leukemia indicate a need for further genetic workup in the context of hematopoietic stem cell transplantation. Blood Cancer J. 2023, 13, 73. [Google Scholar] [CrossRef]
  71. Hirsch, P.; Bories, D.; Chapiro, E.; Nguyen-Khac, F.; Benusiglio, P.R.; Norol, F.; Nguyen, S. Successive relapses from donor and host cells in a patient with DEAD-box helicase 41 (DDX41)-associated myelodysplastic syndrome: The lessons to be learned. Br. J. Haematol. 2022, 199, 623–626. [Google Scholar] [CrossRef] [PubMed]
  72. Huo, L.; Zhang, Z.; Zhou, H.; Xie, J.; Jiang, A.; Wang, Q.; Ding, Z.; Dai, H.; Liu, D.; Wu, N.; et al. Causative germline variant p.Y259C of DDX41 recurrently identified in acute lymphoblastic leukaemia. Br. J. Haematol. 2023, 202, 199–203. [Google Scholar] [CrossRef] [PubMed]
  73. Jelloul, F.Z.; Routbort, M.J.; DiNardo, C.D.; Bueso-Ramos, C.E.; Kanagal-Shamanna, R.; Thakral, B.; Zuo, Z.; Yin, C.C.; Loghavi, S.; Ok, C.Y.; et al. DDX41 mutations in patients with non-myeloid hematologic neoplasms. Am. J. Hematol. 2023, 98, E193–E196. [Google Scholar] [CrossRef] [PubMed]
  74. Ma, J.; Mahmud, N.; Bosland, M.C.; Ross, S.R. DDX41 is needed for pre- and postnatal hematopoietic stem cell differentiation in mice. Stem Cell Rep. 2022, 17, 879–893. [Google Scholar] [CrossRef]
  75. Chlon, T.M.; Stepanchick, E.; Hershberger, C.E.; Daniels, N.J.; Hueneman, K.M.; Kuenzi Davis, A.; Choi, K.; Zheng, Y.; Gurnari, C.; Haferlach, T.; et al. Germline DDX41 mutations cause ineffective hematopoiesis and myelodysplasia. Cell Stem Cell 2021, 28, 1966–1981.e6. [Google Scholar] [CrossRef]
  76. Chen, L.; Chen, J.Y.; Huang, Y.J.; Gu, Y.; Qiu, J.; Qian, H.; Shao, C.; Zhang, X.; Hu, J.; Li, H.; et al. The Augmented R-Loop Is a Unifying Mechanism for Myelodysplastic Syndromes Induced by High-Risk Splicing Factor Mutations. Mol. Cell 2018, 69, 412–425.e6. [Google Scholar] [CrossRef]
  77. Nguyen, H.D.; Leong, W.Y.; Li, W.; Reddy, P.N.G.; Sullivan, J.D.; Walter, M.J.; Zou, L.; Graubert, T.A. Spliceosome Mutations Induce R Loop-Associated Sensitivity to ATR Inhibition in Myelodysplastic Syndromes. Cancer Res. 2018, 78, 5363–5374. [Google Scholar] [CrossRef]
  78. Singh, S.; Ahmed, D.; Dolatshad, H.; Tatwavedi, D.; Schulze, U.; Sanchi, A.; Ryley, S.; Dhir, A.; Carpenter, L.; Watt, S.M.; et al. SF3B1 mutations induce R-loop accumulation and DNA damage in MDS and leukemia cells with therapeutic implications. Leukemia 2020, 34, 2525–2530. [Google Scholar] [CrossRef]
  79. Cusan, M.; Shen, H.; Zhang, B.; Liao, A.; Yang, L.; Jin, M.; Fernandez, M.; Iyer, P.; Wu, Y.; Hart, K.L.; et al. SF3B1 mutation and ATM deletion co-drive leukemogenesis via centromeric R-loop dysregulation. J. Clin. Investig. 2023, 133, e163325. [Google Scholar] [CrossRef]
  80. Weinreb, J.T.; Ghazale, N.; Pradhan, K.; Gupta, V.; Potts, K.S.; Tricomi, B.; Daniels, N.J.; Padgett, R.A.; De Oliveira, S.; Verma, A.; et al. Excessive R-loops trigger an inflammatory cascade leading to increased HSPC production. Dev. Cell 2021, 56, 627–640.e5. [Google Scholar] [CrossRef]
  81. Mosler, T.; Conte, F.; Longo, G.M.C.; Mikicic, I.; Kreim, N.; Möckel, M.M.; Petrosino, G.; Flach, J.; Barau, J.; Luke, B.; et al. R-loop proximity proteomics identifies a role of DDX41 in transcription-associated genomic instability. Nat. Commun. 2021, 12, 7314. [Google Scholar] [CrossRef] [PubMed]
  82. Shinriki, S.; Hirayama, M.; Nagamachi, A.; Yokoyama, A.; Kawamura, T.; Kanai, A.; Kawai, H.; Iwakiri, J.; Liu, R.; Maeshiro, M.; et al. DDX41 coordinates RNA splicing and transcriptional elongation to prevent DNA replication stress in hematopoietic cells. Leukemia 2022, 36, 2605–2620. [Google Scholar] [CrossRef] [PubMed]
  83. Yoshida, K.; Sanada, M.; Shiraishi, Y.; Nowak, D.; Nagata, Y.; Yamamoto, R.; Sato, Y.; Sato-Otsubo, A.; Kon, A.; Nagasaki, M.; et al. Frequent pathway mutations of splicing machinery in myelodysplasia. Nature 2011, 478, 64–69. [Google Scholar] [CrossRef] [PubMed]
  84. Cvitkovic, I.; Jurica, M.S. Spliceosome database: A tool for tracking components of the spliceosome. Nucleic Acids Res. 2013, 41, D132–D141. [Google Scholar] [CrossRef] [PubMed]
  85. Singh, R.S.; Vidhyasagar, V.; Yang, S.; Arna, A.B.; Yadav, M.; Aggarwal, A.; Aguilera, A.N.; Shinriki, S.; Bhanumathy, K.K.; Pandey, K.; et al. DDX41 is required for cGAS-STING activation against DNA virus infection. Cell Rep. 2022, 39, 110856. [Google Scholar] [CrossRef]
  86. Crossley, M.P.; Song, C.; Bocek, M.J.; Choi, J.H.; Kousorous, J.; Sathirachinda, A.; Lin, C.; Brickner, J.R.; Bai, G.; Lans, H.; et al. R-loop-derived cytoplasmic RNA-DNA hybrids activate an immune response. Nature 2023, 613, 187–194. [Google Scholar] [CrossRef]
  87. Challakkara, M.F.; Chhabra, R. snoRNAs in hematopoiesis and blood malignancies: A comprehensive review. J. Cell. Physiol. 2023, 238, 1207–1225. [Google Scholar] [CrossRef]
  88. Dong, J.; Wang, H.; Zhang, Z.; Yang, L.; Qian, X.; Qian, W.; Han, Y.; Huang, H.; Qian, P. Small but strong: Pivotal roles and potential applications of snoRNAs in hematopoietic malignancies. Front. Oncol. 2022, 12, 939465. [Google Scholar] [CrossRef]
  89. Tungalag, S.; Shinriki, S.; Hirayama, M.; Nagamachi, A.; Kanai, A.; Inaba, T.; Matsui, H. Ribosome profiling analysis reveals the roles of DDX41 in translational regulation. Int. J. Hematol. 2023, 117, 876–888. [Google Scholar] [CrossRef]
  90. Ramdzan, Z.M.; Nepveu, A. CUX1, a haploinsufficient tumour suppressor gene overexpressed in advanced cancers. Nat. Rev. Cancer 2014, 14, 673–682. [Google Scholar] [CrossRef]
  91. Imgruet, M.K.; Lutze, J.; An, N.; Hu, B.; Khan, S.; Kurkewich, J.; Martinez, T.C.; Wolfgeher, D.; Gurbuxani, S.K.; Kron, S.J.; et al. Loss of a 7q gene, CUX1, disrupts epigenetically driven DNA repair and drives therapy-related myeloid neoplasms. Blood 2021, 138, 790–805. [Google Scholar] [CrossRef] [PubMed]
  92. Leroy, B.; Anderson, M.; Soussi, T. TP53 mutations in human cancer: Database reassessment and prospects for the next decade. Hum. Mutat. 2014, 35, 672–688. [Google Scholar] [CrossRef] [PubMed]
  93. Hernandez Borrero, L.J.; El-Deiry, W.S. Tumor suppressor p53: Biology, signaling pathways, and therapeutic targeting. Biochim. Biophys. Acta Rev. Cancer 2021, 1876, 188556. [Google Scholar] [CrossRef] [PubMed]
  94. Eisenstein, M. p53: An anticancer protein’s chequered past and promising future. Nature 2022, 603, S1. [Google Scholar] [CrossRef] [PubMed]
  95. Usman, R.M.; Razzaq, F.; Akbar, A.; Farooqui, A.A.; Iftikhar, A.; Latif, A.; Hassan, H.; Zhao, J.; Carew, J.S.; Nawrocki, S.T.; et al. Role and mechanism of autophagy-regulating factors in tumorigenesis and drug resistance. Asia Pac. J. Clin. Oncol. 2021, 17, 193–208. [Google Scholar] [CrossRef] [PubMed]
  96. Mantovani, F.; Collavin, L.; Del Sal, G. Mutant p53 as a guardian of the cancer cell. Cell Death Differ. 2019, 26, 199–212. [Google Scholar] [CrossRef]
  97. Lapke, N.; Lu, Y.J.; Liao, C.T.; Lee, L.Y.; Lin, C.Y.; Wang, H.M.; Ng, S.H.; Chen, S.J.; Yen, T.C. Missense mutations in the TP53 DNA-binding domain predict outcomes in patients with advanced oral cavity squamous cell carcinoma. Oncotarget 2016, 7, 44194–44210. [Google Scholar] [CrossRef]
  98. Hansen, S.; Hupp, T.R.; Lane, D.P. Allosteric regulation of the thermostability and DNA binding activity of human p53 by specific interacting proteins. CRC Cell Transformation Group. J. Biol. Chem. 1996, 271, 3917–3924. [Google Scholar] [CrossRef]
  99. Alvarado-Ortiz, E.; de la Cruz-Lopez, K.G.; Becerril-Rico, J.; Sarabia-Sanchez, M.A.; Ortiz-Sanchez, E.; Garcia-Carranca, A. Mutant p53 Gain-of-Function: Role in Cancer Development, Progression, and Therapeutic Approaches. Front. Cell Dev. Biol. 2020, 8, 607670. [Google Scholar] [CrossRef]
  100. Gencel-Augusto, J.; Lozano, G. p53 tetramerization: At the center of the dominant-negative effect of mutant p53. Genes Dev. 2020, 34, 1128–1146. [Google Scholar] [CrossRef]
  101. Zhu, G.; Pan, C.; Bei, J.X.; Li, B.; Liang, C.; Xu, Y.; Fu, X. Mutant p53 in Cancer Progression and Targeted Therapies. Front. Oncol. 2020, 10, 595187. [Google Scholar] [CrossRef] [PubMed]
  102. Keymling, M.; Schlemmer, H.P.; Kratz, C.; Pfeil, A.; Bickelhaupt, S.; Alsady, T.M.; Renz, D.M. Li-Fraumeni syndrome. Radiologie 2022, 62, 1026–1032. [Google Scholar] [CrossRef] [PubMed]
  103. Sejben, A.; Tiszlavicz, L.; Polyak, K.; Kovacs, L.; Maraz, A.; Torok, D.; Lepran, A.; Ottlakan, A.; Furak, J. Li-Fraumeni syndrome. Orvosi Hetil. 2019, 160, 228–234. [Google Scholar] [CrossRef] [PubMed]
  104. Qian, M.; Cao, X.; Devidas, M.; Yang, W.; Cheng, C.; Dai, Y.; Carroll, A.; Heerema, N.A.; Zhang, H.; Moriyama, T.; et al. TP53 Germline Variations Influence the Predisposition and Prognosis of B-Cell Acute Lymphoblastic Leukemia in Children. J. Clin. Oncol. 2018, 36, 591–599. [Google Scholar] [CrossRef] [PubMed]
  105. Comeaux, E.Q.; Mullighan, C.G. TP53 Mutations in Hypodiploid Acute Lymphoblastic Leukemia. Cold Spring Harb. Perspect. Med. 2017, 7, a026286. [Google Scholar] [CrossRef]
  106. Preudhomme, C.; Sagot, C.; Boissel, N.; Cayuela, J.M.; Tigaud, I.; de Botton, S.; Thomas, X.; Raffoux, E.; Lamandin, C.; Castaigne, S.; et al. Favorable prognostic significance of CEBPA mutations in patients with de novo acute myeloid leukemia: A study from the Acute Leukemia French Association (ALFA). Blood 2002, 100, 2717–2723. [Google Scholar] [CrossRef]
  107. Avellino, R.; Delwel, R. Expression and regulation of C/EBPalpha in normal myelopoiesis and in malignant transformation. Blood 2017, 129, 2083–2091. [Google Scholar] [CrossRef]
  108. Pabst, T.; Mueller, B.U. Complexity of CEBPA dysregulation in human acute myeloid leukemia. Clin. Cancer Res. 2009, 15, 5303–5307. [Google Scholar] [CrossRef]
  109. Tsukada, J.; Yoshida, Y.; Kominato, Y.; Auron, P.E. The CCAAT/enhancer (C/EBP) family of basic-leucine zipper (bZIP) transcription factors is a multifaceted highly-regulated system for gene regulation. Cytokine 2011, 54, 6–19. [Google Scholar] [CrossRef]
  110. Pulikkan, J.A.; Tenen, D.G.; Behre, G. C/EBPalpha deregulation as a paradigm for leukemogenesis. Leukemia 2017, 31, 2279–2285. [Google Scholar] [CrossRef]
  111. Tawana, K.; Fitzgibbon, J. CEBPA-Associated Familial Acute Myeloid Leukemia (AML). In GeneReviews®; Adam, M.P., Mirzaa, G.M., Pagon, R.A., Wallace, S.E., Bean, L.J.H., Gripp, K.W., Amemiya, A., Eds.; University of Washington: Seattle, WA, USA, 1993. [Google Scholar]
  112. West, A.H.; Godley, L.A.; Churpek, J.E. Familial myelodysplastic syndrome/acute leukemia syndromes: A review and utility for translational investigations. Ann. N. Y Acad. Sci. 2014, 1310, 111–118. [Google Scholar] [CrossRef] [PubMed]
  113. Godley, L.A. Inherited predisposition to acute myeloid leukemia. Semin. Hematol. 2014, 51, 306–321. [Google Scholar] [CrossRef] [PubMed]
  114. Pabst, T.; Mueller, B.U.; Zhang, P.; Radomska, H.S.; Narravula, S.; Schnittger, S.; Behre, G.; Hiddemann, W.; Tenen, D.G. Dominant-negative mutations of CEBPA, encoding CCAAT/enhancer binding protein-alpha (C/EBPalpha), in acute myeloid leukemia. Nat. Genet. 2001, 27, 263–270. [Google Scholar] [CrossRef] [PubMed]
  115. Frohling, S.; Schlenk, R.F.; Stolze, I.; Bihlmayr, J.; Benner, A.; Kreitmeier, S.; Tobis, K.; Dohner, H.; Dohner, K. CEBPA mutations in younger adults with acute myeloid leukemia and normal cytogenetics: Prognostic relevance and analysis of cooperating mutations. J. Clin. Oncol. 2004, 22, 624–633. [Google Scholar] [CrossRef]
  116. Xiao, H.; Shi, J.; Luo, Y.; Tan, Y.; He, J.; Xie, W.; Zhang, L.; Wang, Y.; Liu, L.; Wu, K.; et al. First report of multiple CEBPA mutations contributing to donor origin of leukemia relapse after allogeneic hematopoietic stem cell transplantation. Blood 2011, 117, 5257–5260. [Google Scholar] [CrossRef]
  117. Vitor, A.C.; Huertas, P.; Legube, G.; de Almeida, S.F. Studying DNA Double-Strand Break Repair: An Ever-Growing Toolbox. Front. Mol. Biosci. 2020, 7, 24. [Google Scholar] [CrossRef]
  118. Gelot, C.; Magdalou, I.; Lopez, B.S. Replication stress in Mammalian cells and its consequences for mitosis. Genes 2015, 6, 267–298. [Google Scholar] [CrossRef]
  119. Arnoult, N.; Correia, A.; Ma, J.; Merlo, A.; Garcia-Gomez, S.; Maric, M.; Tognetti, M.; Benner, C.W.; Boulton, S.J.; Saghatelian, A.; et al. Regulation of DNA repair pathway choice in S and G2 phases by the NHEJ inhibitor CYREN. Nature 2017, 549, 548–552. [Google Scholar] [CrossRef]
  120. Tung, N.; Battelli, C.; Allen, B.; Kaldate, R.; Bhatnagar, S.; Bowles, K.; Timms, K.; Garber, J.E.; Herold, C.; Ellisen, L.; et al. Frequency of mutations in individuals with breast cancer referred for BRCA1 and BRCA2 testing using next-generation sequencing with a 25-gene panel. Cancer 2015, 121, 25–33. [Google Scholar] [CrossRef]
  121. Mateo, J.; Carreira, S.; Sandhu, S.; Miranda, S.; Mossop, H.; Perez-Lopez, R.; Nava Rodrigues, D.; Robinson, D.; Omlin, A.; Tunariu, N.; et al. DNA-Repair Defects and Olaparib in Metastatic Prostate Cancer. N. Engl. J. Med. 2015, 373, 1697–1708. [Google Scholar] [CrossRef]
  122. Messina, C.; Cattrini, C.; Soldato, D.; Vallome, G.; Caffo, O.; Castro, E.; Olmos, D.; Boccardo, F.; Zanardi, E. BRCA Mutations in Prostate Cancer: Prognostic and Predictive Implications. J. Oncol. 2020, 2020, 4986365. [Google Scholar] [CrossRef] [PubMed]
  123. Goggins, M.; Overbeek, K.A.; Brand, R.; Syngal, S.; Del Chiaro, M.; Bartsch, D.K.; Bassi, C.; Carrato, A.; Farrell, J.; Fishman, E.K.; et al. Management of patients with increased risk for familial pancreatic cancer: Updated recommendations from the International Cancer of the Pancreas Screening (CAPS) Consortium. Gut 2020, 69, 7–17. [Google Scholar] [CrossRef] [PubMed]
  124. Stolarova, L.; Kleiblova, P.; Janatova, M.; Soukupova, J.; Zemankova, P.; Macurek, L.; Kleibl, Z. CHEK2 Germline Variants in Cancer Predisposition: Stalemate Rather than Checkmate. Cells 2020, 9, 2675. [Google Scholar] [CrossRef] [PubMed]
  125. Fostira, F.; Kostantopoulou, I.; Apostolou, P.; Papamentzelopoulou, M.S.; Papadimitriou, C.; Faliakou, E.; Christodoulou, C.; Boukovinas, I.; Razis, E.; Tryfonopoulos, D.; et al. One in three highly selected Greek patients with breast cancer carries a loss-of-function variant in a cancer susceptibility gene. J. Med. Genet. 2020, 57, 53–61. [Google Scholar] [CrossRef] [PubMed]
  126. Pritchard, C.C.; Mateo, J.; Walsh, M.F.; De Sarkar, N.; Abida, W.; Beltran, H.; Garofalo, A.; Gulati, R.; Carreira, S.; Eeles, R.; et al. Inherited DNA-Repair Gene Mutations in Men with Metastatic Prostate Cancer. N. Engl. J. Med. 2016, 375, 443–453. [Google Scholar] [CrossRef] [PubMed]
  127. Carlo, M.I.; Mukherjee, S.; Mandelker, D.; Vijai, J.; Kemel, Y.; Zhang, L.; Knezevic, A.; Patil, S.; Ceyhan-Birsoy, O.; Huang, K.C.; et al. Prevalence of Germline Mutations in Cancer Susceptibility Genes in Patients with Advanced Renal Cell Carcinoma. JAMA Oncol. 2018, 4, 1228–1235. [Google Scholar] [CrossRef]
  128. Kaczmarek-Rys, M.; Ziemnicka, K.; Hryhorowicz, S.T.; Gorczak, K.; Hoppe-Golebiewska, J.; Skrzypczak-Zielinska, M.; Tomys, M.; Golab, M.; Szkudlarek, M.; Budny, B.; et al. The c.470 T > C CHEK2 missense variant increases the risk of differentiated thyroid carcinoma in the Great Poland population. Hered. Cancer Clin. Pract. 2015, 13, 8. [Google Scholar] [CrossRef]
  129. Katona, B.W.; Yang, Y.X. Colorectal cancer risk associated with the CHEK2 1100delC variant. Eur. J. Cancer 2017, 83, 103–105. [Google Scholar] [CrossRef]
  130. Laitman, Y.; Nielsen, S.M.; Hatchell, K.E.; Truty, R.; Bernstein-Molho, R.; Esplin, E.D.; Friedman, E. Re-evaluating cancer risks associated with the CHEK2 p.Ser428Phe Ashkenazi Jewish founder pathogenic variant. Fam. Cancer 2022, 21, 305–308. [Google Scholar] [CrossRef]
  131. Janiszewska, H.; Bak, A.; Pilarska, M.; Heise, M.; Junkiert-Czarnecka, A.; Kuliszkiewicz-Janus, M.; Calbecka, M.; Jazwiec, B.; Wolowiec, D.; Kuliczkowski, K.; et al. A risk of essential thrombocythemia in carriers of constitutional CHEK2 gene mutations. Haematologica 2012, 97, 366–370. [Google Scholar] [CrossRef]
  132. Bao, E.L.; Nandakumar, S.K.; Liao, X.; Bick, A.G.; Karjalainen, J.; Tabaka, M.; Gan, O.I.; Havulinna, A.S.; Kiiskinen, T.T.J.; Lareau, C.A.; et al. Inherited myeloproliferative neoplasm risk affects haematopoietic stem cells. Nature 2020, 586, 769–775. [Google Scholar] [CrossRef] [PubMed]
  133. Janiszewska, H.; Bak, A.; Hartwig, M.; Kuliszkiewicz-Janus, M.; Calbecka, M.; Jazwiec, B.; Kuliczkowski, K.; Haus, O. The germline mutations of the CHEK2 gene are associated with an increased risk of polycythaemia vera. Br. J. Haematol. 2016, 173, 150–152. [Google Scholar] [CrossRef] [PubMed]
  134. Janiszewska, H.; Bak, A.; Skonieczka, K.; Jaskowiec, A.; Kielbinski, M.; Jachalska, A.; Czyzewska, M.; Jazwiec, B.; Kuliszkiewicz-Janus, M.; Czyz, J.; et al. Constitutional mutations of the CHEK2 gene are a risk factor for MDS, but not for de novo AML. Leuk. Res. 2018, 70, 74–78. [Google Scholar] [CrossRef] [PubMed]
  135. Sharifi, M.J.; Zaker, F.; Nasiri, N.; Yaghmaie, M. Epigenetic changes in FOXO3 and CHEK2 genes and their correlation with clinicopathological findings in myelodysplastic syndromes. Hematol. Oncol. Stem Cell Ther. 2020, 13, 214–219. [Google Scholar] [CrossRef] [PubMed]
  136. Berger, G.; van den Berg, E.; Smetsers, S.; Leegte, B.K.; Sijmons, R.H.; Abbott, K.M.; Mulder, A.B.; Vellenga, E. Fanconi anaemia presenting as acute myeloid leukaemia and myelodysplastic syndrome in adulthood: A family report on co-occurring FANCC and CHEK2 mutations. Br. J. Haematol. 2019, 184, 1071–1073. [Google Scholar] [CrossRef]
  137. Paperna, T.; Sharon-Shwartzman, N.; Kurolap, A.; Goldberg, Y.; Moustafa, N.; Carasso, Y.; Feinstien, M.; Mory, A.; Reznick-Levi, G.; Gonzaga-Jauregui, C.; et al. Homozygosity for CHEK2 p.Gly167Arg leads to a unique cancer syndrome with multiple complex chromosomal translocations in peripheral blood karyotype. J. Med. Genet. 2020, 57, 500–504. [Google Scholar] [CrossRef] [PubMed]
  138. Hofmann, W.K.; Miller, C.W.; Tsukasaki, K.; Tavor, S.; Ikezoe, T.; Hoelzer, D.; Takeuchi, S.; Koeffler, H.P. Mutation analysis of the DNA-damage checkpoint gene CHK2 in myelodysplastic syndromes and acute myeloid leukemias. Leuk. Res. 2001, 25, 333–338. [Google Scholar] [CrossRef] [PubMed]
  139. Ye, Q.; Chen, H.; Wen, Z.; Guo, W.; Huang, Y.; Mo, X. Abnormal expression of p-ATM/CHK2 in nasal extranodal NK/T cell lymphoma, nasal type, is correlated with poor prognosis. J. Clin. Pathol. 2021, 74, 223–227. [Google Scholar] [CrossRef]
  140. Ueno, S.; Sudo, T.; Hirasawa, A. ATM: Functions of ATM Kinase and Its Relevance to Hereditary Tumors. Int. J. Mol. Sci. 2022, 23, 523. [Google Scholar] [CrossRef]
  141. Nakajima, H. Genetic abnormalities in AML. Rinsho Ketsueki 2019, 60, 584–593. [Google Scholar] [CrossRef]
  142. Rothblum-Oviatt, C.; Wright, J.; Lefton-Greif, M.A.; McGrath-Morrow, S.A.; Crawford, T.O.; Lederman, H.M. Ataxia telangiectasia: A review. Orphanet J. Rare Dis. 2016, 11, 159. [Google Scholar] [CrossRef] [PubMed]
  143. Swift, M.; Chase, C.L.; Morrell, D. Cancer predisposition of ataxia-telangiectasia heterozygotes. Cancer Genet. Cytogenet. 1990, 46, 21–27. [Google Scholar] [CrossRef] [PubMed]
  144. Ito, K.; Hirao, A.; Arai, F.; Matsuoka, S.; Takubo, K.; Hamaguchi, I.; Nomiyama, K.; Hosokawa, K.; Sakurada, K.; Nakagata, N.; et al. Regulation of oxidative stress by ATM is required for self-renewal of haematopoietic stem cells. Nature 2004, 431, 997–1002. [Google Scholar] [CrossRef] [PubMed]
  145. Gregory, M.A.; D’Alessandro, A.; Alvarez-Calderon, F.; Kim, J.; Nemkov, T.; Adane, B.; Rozhok, A.I.; Kumar, A.; Kumar, V.; Pollyea, D.A.; et al. ATM/G6PD-driven redox metabolism promotes FLT3 inhibitor resistance in acute myeloid leukemia. Proc. Natl. Acad. Sci. USA 2016, 113, E6669–E6678. [Google Scholar] [CrossRef] [PubMed]
  146. Liu, X.; Liao, W.; Peng, H.; Luo, X.; Luo, Z.; Jiang, H.; Xu, L. miR-181a promotes G1/S transition and cell proliferation in pediatric acute myeloid leukemia by targeting ATM. J. Cancer Res. Clin. Oncol. 2016, 142, 77–87. [Google Scholar] [CrossRef] [PubMed]
  147. Schrader, A.; Crispatzu, G.; Oberbeck, S.; Mayer, P.; Putzer, S.; von Jan, J.; Vasyutina, E.; Warner, K.; Weit, N.; Pflug, N.; et al. Actionable perturbations of damage responses by TCL1/ATM and epigenetic lesions form the basis of T-PLL. Nat. Commun. 2018, 9, 697. [Google Scholar] [CrossRef] [PubMed]
  148. Wang, X.; Song, Y.; Chen, W.; Ding, N.; Liu, W.; Xie, Y.; Wang, Y.; Zhu, J.; Zeng, C. Germline variants of DNA repair genes in early onset mantle cell lymphoma. Oncogene 2021, 40, 551–563. [Google Scholar] [CrossRef]
  149. Tiao, G.; Improgo, M.R.; Kasar, S.; Poh, W.; Kamburov, A.; Landau, D.A.; Tausch, E.; Taylor-Weiner, A.; Cibulskis, C.; Bahl, S.; et al. Rare germline variants in ATM are associated with chronic lymphocytic leukemia. Leukemia 2017, 31, 2244–2247. [Google Scholar] [CrossRef]
  150. Homan, C.C.; Scott, H.S.; Brown, A.L. Hereditary platelet disorders associated with germ line variants in RUNX1, ETV6, and ANKRD26. Blood 2023, 141, 1533–1543. [Google Scholar] [CrossRef]
  151. Galera, P.; Dulau-Florea, A.; Calvo, K.R. Inherited thrombocytopenia and platelet disorders with germline predisposition to myeloid neoplasia. Int. J. Lab. Hematol. 2019, 41 (Suppl. 1), 131–141. [Google Scholar] [CrossRef] [PubMed]
  152. Asou, N. The role of a Runt domain transcription factor AML1/RUNX1 in leukemogenesis and its clinical implications. Crit. Rev. Oncol. Hematol. 2003, 45, 129–150. [Google Scholar] [CrossRef] [PubMed]
  153. Okumura, A.J.; Peterson, L.F.; Okumura, F.; Boyapati, A.; Zhang, D.E. t(8;21)(q22;q22) Fusion proteins preferentially bind to duplicated AML1/RUNX1 DNA-binding sequences to differentially regulate gene expression. Blood 2008, 112, 1392–1401. [Google Scholar] [CrossRef] [PubMed]
  154. Kamath-Loeb, A.S.; Shen, J.C.; Schmitt, M.W.; Kohrn, B.F.; Loeb, K.R.; Estey, E.H.; Dai, J.; Chien, S.; Loeb, L.A.; Becker, P.S. Accurate detection of subclonal variants in paired diagnosis-relapse acute myeloid leukemia samples by next generation Duplex Sequencing. Leuk. Res. 2022, 115, 106822. [Google Scholar] [CrossRef] [PubMed]
  155. Zharlyganova, D.; Harada, H.; Harada, Y.; Shinkarev, S.; Zhumadilov, Z.; Zhunusova, A.; Tchaizhunusova, N.J.; Apsalikov, K.N.; Kemaikin, V.; Zhumadilov, K.; et al. High frequency of AML1/RUNX1 point mutations in radiation-associated myelodysplastic syndrome around Semipalatinsk nuclear test site. J. Radiat. Res. 2008, 49, 549–555. [Google Scholar] [CrossRef] [PubMed]
  156. Sendker, S.; Awada, A.; Domagalla, S.; Sendker, M.; Orhan, E.; Hoffmeister, L.M.; Antoniou, E.; Niktoreh, N.; Reinhardt, D.; von Neuhoff, N.; et al. RUNX1 mutation has no prognostic significance in paediatric AML: A retrospective study of the AML-BFM study group. Leukemia 2023, 37, 1435–1443. [Google Scholar] [CrossRef] [PubMed]
  157. Sood, R.; Kamikubo, Y.; Liu, P. Role of RUNX1 in hematological malignancies. Blood 2017, 129, 2070–2082. [Google Scholar] [CrossRef] [PubMed]
  158. Hayashi, Y.; Harada, Y.; Huang, G.; Harada, H. Myeloid neoplasms with germ line RUNX1 mutation. Int. J. Hematol. 2017, 106, 183–188. [Google Scholar] [CrossRef]
  159. Bellissimo, D.C.; Speck, N.A. RUNX1 Mutations in Inherited and Sporadic Leukemia. Front. Cell Dev. Biol. 2017, 5, 111. [Google Scholar] [CrossRef]
  160. Ng, I.K.; Lee, J.; Ng, C.; Kosmo, B.; Chiu, L.; Seah, E.; Mok, M.M.H.; Tan, K.; Osato, M.; Chng, W.J.; et al. Preleukemic and second-hit mutational events in an acute myeloid leukemia patient with a novel germline RUNX1 mutation. Biomark. Res. 2018, 6, 16. [Google Scholar] [CrossRef]
  161. Hong, D.; Fritz, A.J.; Gordon, J.A.; Tye, C.E.; Boyd, J.R.; Tracy, K.M.; Frietze, S.E.; Carr, F.E.; Nickerson, J.A.; Van Wijnen, A.J.; et al. RUNX1-dependent mechanisms in biological control and dysregulation in cancer. J. Cell. Physiol. 2019, 234, 8597–8609. [Google Scholar] [CrossRef]
  162. Deuitch, N.; Broadbridge, E.; Cunningham, L.; Liu, P. RUNX1 Familial Platelet Disorder with Associated Myeloid Malignancies. In GeneReviews®; Adam, M.P., Mirzaa, G.M., Pagon, R.A., Wallace, S.E., Bean, L.J.H., Gripp, K.W., Amemiya, A., Eds.; University of Washington: Seattle, WA, USA, 1993. [Google Scholar]
  163. Homan, C.C.; Drazer, M.W.; Yu, K.; Lawrence, D.M.; Feng, J.; Arriola-Martinez, L.; Pozsgai, M.J.; McNeely, K.E.; Ha, T.; Venugopal, P.; et al. Somatic mutational landscape of hereditary hematopoietic malignancies caused by germline variants in RUNX1, GATA2, and DDX41. Blood Adv. 2023, 7, 6092–6107. [Google Scholar] [CrossRef] [PubMed]
  164. Vyas, H.; Alcheikh, A.; Lowe, G.; Stevenson, W.S.; Morgan, N.V.; Rabbolini, D.J. Prevalence and natural history of variants in the ANKRD26 gene: A short review and update of reported cases. Platelets 2022, 33, 1107–1112. [Google Scholar] [CrossRef] [PubMed]
  165. Ferrari, S.; Regazzo, D.; Omenetto, E.; Scaroni, C.; Semenzato, G.; Fabris, F.; Vianello, F. A novel RUNX1 mutation with ANKRD26 dysregulation is related to thrombocytopenia in a sporadic form of myelodysplastic syndrome. Aging Clin. Exp. Res. 2021, 33, 1987–1992. [Google Scholar] [CrossRef] [PubMed]
  166. Sullivan, M.J.; Palmer, E.L.; Botero, J.P. ANKRD26-Related Thrombocytopenia and Predisposition to Myeloid Neoplasms. Curr. Hematol. Malig. Rep. 2022, 17, 105–112. [Google Scholar] [CrossRef] [PubMed]
  167. Kennedy, A.L.; Shimamura, A. Genetic predisposition to MDS: Clinical features and clonal evolution. Blood 2019, 133, 1071–1085. [Google Scholar] [CrossRef] [PubMed]
  168. Melazzini, F.; Palombo, F.; Balduini, A.; De Rocco, D.; Marconi, C.; Noris, P.; Gnan, C.; Pippucci, T.; Bozzi, V.; Faleschini, M.; et al. Clinical and pathogenic features of ETV6-related thrombocytopenia with predisposition to acute lymphoblastic leukemia. Haematologica 2016, 101, 1333–1342. [Google Scholar] [CrossRef]
  169. Wang, Q.; Dong, S.; Yao, H.; Wen, L.; Qiu, H.; Qin, L.; Ma, L.; Chen, S. ETV6 mutation in a cohort of 970 patients with hematologic malignancies. Haematologica 2014, 99, e176–e178. [Google Scholar] [CrossRef]
  170. Di Paola, J.; Porter, C.C. ETV6-related thrombocytopenia and leukemia predisposition. Blood 2019, 134, 663–667. [Google Scholar] [CrossRef]
  171. Rodriguez-Hernandez, G.; Casado-Garcia, A.; Isidro-Hernandez, M.; Picard, D.; Raboso-Gallego, J.; Aleman-Arteaga, S.; Orfao, A.; Blanco, O.; Riesco, S.; Prieto-Matos, P.; et al. The Second Oncogenic Hit Determines the Cell Fate of ETV6-RUNX1 Positive Leukemia. Front. Cell Dev. Biol. 2021, 9, 704591. [Google Scholar] [CrossRef]
  172. Filipiuk, A.; Kozakiewicz, A.; Kosmider, K.; Lejman, M.; Zawitkowska, J. Genetic Disorders with Predisposition to Paediatric Haematopoietic Malignancies-A Review. Cancers 2022, 14, 3569. [Google Scholar] [CrossRef]
  173. Schafer, D.; Olsen, M.; Lahnemann, D.; Stanulla, M.; Slany, R.; Schmiegelow, K.; Borkhardt, A.; Fischer, U. Five percent of healthy newborns have an ETV6-RUNX1 fusion as revealed by DNA-based GIPFEL screening. Blood 2018, 131, 821–826. [Google Scholar] [CrossRef] [PubMed]
  174. Tesi, B.; Davidsson, J.; Voss, M.; Rahikkala, E.; Holmes, T.D.; Chiang, S.C.C.; Komulainen-Ebrahim, J.; Gorcenco, S.; Rundberg Nilsson, A.; Ripperger, T.; et al. Gain-of-function SAMD9L mutations cause a syndrome of cytopenia, immunodeficiency, MDS, and neurological symptoms. Blood 2017, 129, 2266–2279. [Google Scholar] [CrossRef] [PubMed]
  175. Pastor, V.B.; Sahoo, S.S.; Boklan, J.; Schwabe, G.C.; Saribeyoglu, E.; Strahm, B.; Lebrecht, D.; Voss, M.; Bryceson, Y.T.; Erlacher, M.; et al. Constitutional SAMD9L mutations cause familial myelodysplastic syndrome and transient monosomy 7. Haematologica 2018, 103, 427–437. [Google Scholar] [CrossRef] [PubMed]
  176. Mekhedov, S.L.; Makarova, K.S.; Koonin, E.V. The complex domain architecture of SAMD9 family proteins, predicted STAND-like NTPases, suggests new links to inflammation and apoptosis. Biol. Direct 2017, 12, 13. [Google Scholar] [CrossRef]
  177. Nagamachi, A.; Matsui, H.; Asou, H.; Ozaki, Y.; Aki, D.; Kanai, A.; Takubo, K.; Suda, T.; Nakamura, T.; Wolff, L.; et al. Haploinsufficiency of SAMD9L, an endosome fusion facilitator, causes myeloid malignancies in mice mimicking human diseases with monosomy 7. Cancer Cell 2013, 24, 305–317. [Google Scholar] [CrossRef] [PubMed]
  178. Nagamachi, A.; Kanai, A.; Nakamura, M.; Okuda, H.; Yokoyama, A.; Shinriki, S.; Matsui, H.; Inaba, T. Multiorgan failure with abnormal receptor metabolism in mice mimicking Samd9/9L syndromes. J. Clin. Investig. 2021, 131, e140147. [Google Scholar] [CrossRef]
  179. Meng, X.; Xiang, Y. RNA granules associated with SAMD9-mediated poxvirus restriction are similar to antiviral granules in composition but do not require TIA1 for poxvirus restriction. Virology 2019, 529, 16–22. [Google Scholar] [CrossRef]
  180. Zhang, F.; Ji, Q.; Chaturvedi, J.; Morales, M.; Mao, Y.; Meng, X.; Dong, L.; Deng, J.; Qian, S.B.; Xiang, Y. Human SAMD9 is a poxvirus-activatable anticodon nuclease inhibiting codon-specific protein synthesis. Sci. Adv. 2023, 9, eadh8502. [Google Scholar] [CrossRef]
  181. Nagata, Y.; Narumi, S.; Guan, Y.; Przychodzen, B.P.; Hirsch, C.M.; Makishima, H.; Shima, H.; Aly, M.; Pastor, V.; Kuzmanovic, T.; et al. Germline loss-of-function SAMD9 and SAMD9L alterations in adult myelodysplastic syndromes. Blood 2018, 132, 2309–2313. [Google Scholar] [CrossRef]
  182. Yahata, T.; Takanashi, T.; Muguruma, Y.; Ibrahim, A.A.; Matsuzawa, H.; Uno, T.; Sheng, Y.; Onizuka, M.; Ito, M.; Kato, S.; et al. Accumulation of oxidative DNA damage restricts the self-renewal capacity of human hematopoietic stem cells. Blood 2011, 118, 2941–2950. [Google Scholar] [CrossRef]
  183. Zhou, T.; Hasty, P.; Walter, C.A.; Bishop, A.J.; Scott, L.M.; Rebel, V.I. Myelodysplastic syndrome: An inability to appropriately respond to damaged DNA? Exp. Hematol. 2013, 41, 665–674. [Google Scholar] [CrossRef] [PubMed]
  184. Thomas, M.E., 3rd; Abdelhamed, S.; Hiltenbrand, R.; Schwartz, J.R.; Sakurada, S.M.; Walsh, M.; Song, G.; Ma, J.; Pruett-Miller, S.M.; Klco, J.M. Pediatric MDS and bone marrow failure-associated germline mutations in SAMD9 and SAMD9L impair multiple pathways in primary hematopoietic cells. Leukemia 2021, 35, 3232–3244. [Google Scholar] [CrossRef] [PubMed]
  185. Milyavsky, M.; Gan, O.I.; Trottier, M.; Komosa, M.; Tabach, O.; Notta, F.; Lechman, E.; Hermans, K.G.; Eppert, K.; Konovalova, Z.; et al. A distinctive DNA damage response in human hematopoietic stem cells reveals an apoptosis-independent role for p53 in self-renewal. Cell Stem Cell 2010, 7, 186–197. [Google Scholar] [CrossRef] [PubMed]
  186. Parker, J.E.; Mufti, G.J.; Rasool, F.; Mijovic, A.; Devereux, S.; Pagliuca, A. The role of apoptosis, proliferation, and the Bcl-2-related proteins in the myelodysplastic syndromes and acute myeloid leukemia secondary to MDS. Blood 2000, 96, 3932–3938. [Google Scholar] [CrossRef] [PubMed]
  187. Tanase-Nakao, K.; Olson, T.S.; Narumi, S. MIRAGE Syndrome. In GeneReviews®; Adam, M.P., Mirzaa, G.M., Pagon, R.A., Wallace, S.E., Bean, L.J.H., Gripp, K.W., Amemiya, A., Eds.; University of Washington: Seattle, WA, USA, 1993. [Google Scholar]
  188. Yoshizaki, K.; Hachiya, R.; Tomobe, Y.; Kaku, U.; Akiba, K.; Shima, H.; Narumi, S.; Hasegawa, Y. MIRAGE syndrome with recurrent pneumonia probably associated with gastroesophageal reflux and achalasia: A case report. Clin. Pediatr. Endocrinol. 2019, 28, 147–153. [Google Scholar] [CrossRef] [PubMed]
  189. Viaene, A.N.; Harding, B.N. The Neuropathology of MIRAGE Syndrome. J. Neuropathol. Exp. Neurol. 2020, 79, 458–462. [Google Scholar] [CrossRef]
  190. Basilious, A.; Basilious, A.; ElJalbout, R.; Robert, M.C. Lacrimal Gland Hypoplasia and Corneal Anesthesia in MIRAGE Syndrome: A Case Report and Literature Review. Cornea 2022, 41, 1041–1044. [Google Scholar] [CrossRef]
  191. Janjua, D.; Shankar, S.; AlMaazmi, M.; Jadhav, D.V. MIRAGE Syndrome Enteropathy Responding to Pancrelipase Despite Normal Pancreatic Fecal Elastase: A Case Report. Am. J. Case Rep. 2022, 23, e937057. [Google Scholar] [CrossRef]
  192. Gorcenco, S.; Komulainen-Ebrahim, J.; Nordborg, K.; Suo-Palosaari, M.; Andreasson, S.; Kruger, J.; Nilsson, C.; Kjellstrom, U.; Rahikkala, E.; Turkiewicz, D.; et al. Ataxia-pancytopenia syndrome with SAMD9L mutations. Neurol. Genet. 2017, 3, e183. [Google Scholar] [CrossRef]
  193. Vaughan, D.; Bogdanova-Mihaylova, P.; Costello, D.J.; Sweeney, B.J.; McNamara, B.; Walsh, R.A.; Murphy, S.M. Ataxia pancytopenia syndrome due to SAMD9L mutation presenting as demyelinating neuropathy. J. Peripher. Nerv. Syst. 2020, 25, 433–437. [Google Scholar] [CrossRef]
  194. King-Robson, J.; Marshall, J.; Smith, F.; Willoughby, L.; Mansour, S.; Sztriha, L. Ataxia-Pancytopenia Syndrome due to a de Novo SAMD9L Mutation. Neurol. Genet. 2021, 7, e580. [Google Scholar] [CrossRef] [PubMed]
  195. Raskind, W.H.; Chen, D.H.; Bird, T. SAMD9L Ataxia-Pancytopenia Syndrome. In GeneReviews®; Adam, M.P., Mirzaa, G.M., Pagon, R.A., Wallace, S.E., Bean, L.J.H., Gripp, K.W., Amemiya, A., Eds.; University of Washington: Seattle, WA, USA, 1993. [Google Scholar]
  196. Smeenk, L.; Ottema, S.; Mulet-Lazaro, R.; Ebert, A.; Havermans, M.; Varea, A.A.; Fellner, M.; Pastoors, D.; van Herk, S.; Erpelinck-Verschueren, C.; et al. Selective Requirement of MYB for Oncogenic Hyperactivation of a Translocated Enhancer in Leukemia. Cancer Discov. 2021, 11, 2868–2883. [Google Scholar] [CrossRef] [PubMed]
  197. Belloucif, Y.; Lobry, C. Super-Enhancers Dysregulations in Hematological Malignancies. Cells 2022, 11, 196. [Google Scholar] [CrossRef] [PubMed]
  198. de Pater, E.; Kaimakis, P.; Vink, C.S.; Yokomizo, T.; Yamada-Inagawa, T.; van der Linden, R.; Kartalaei, P.S.; Camper, S.A.; Speck, N.; Dzierzak, E. Gata2 is required for HSC generation and survival. J. Exp. Med. 2013, 210, 2843–2850. [Google Scholar] [CrossRef] [PubMed]
  199. Santiago, M.; Liquori, A.; Such, E.; Zuniga, A.; Cervera, J. The Clinical Spectrum, Diagnosis, and Management of GATA2 Deficiency. Cancers 2023, 15, 1590. [Google Scholar] [CrossRef]
  200. Wehr, C.; Grotius, K.; Casadei, S.; Bleckmann, D.; Bode, S.F.N.; Frye, B.C.; Seidl, M.; Gulsuner, S.; King, M.C.; Percival, M.B.; et al. A novel disease-causing synonymous exonic mutation in GATA2 affecting RNA splicing. Blood 2018, 132, 1211–1215. [Google Scholar] [CrossRef] [PubMed]
  201. Oleaga-Quintas, C.; de Oliveira-Junior, E.B.; Rosain, J.; Rapaport, F.; Deswarte, C.; Guerin, A.; Sajjath, S.M.; Zhou, Y.J.; Marot, S.; Lozano, C.; et al. Inherited GATA2 Deficiency Is Dominant by Haploinsufficiency and Displays Incomplete Clinical Penetrance. J. Clin. Immunol. 2021, 41, 639–657. [Google Scholar] [CrossRef]
  202. Calvo, K.R.; Hickstein, D.D. The spectrum of GATA2 deficiency syndrome. Blood 2023, 141, 1524–1532. [Google Scholar] [CrossRef]
  203. McReynolds, L.J.; Calvo, K.R.; Holland, S.M. Germline GATA2 Mutation and Bone Marrow Failure. Hematol. Oncol. Clin. N. Am. 2018, 32, 713–728. [Google Scholar] [CrossRef]
  204. Mir, M.A.; Kochuparambil, S.T.; Abraham, R.S.; Rodriguez, V.; Howard, M.; Hsu, A.P.; Jackson, A.E.; Holland, S.M.; Patnaik, M.M. Spectrum of myeloid neoplasms and immune deficiency associated with germline GATA2 mutations. Cancer Med. 2015, 4, 490–499. [Google Scholar] [CrossRef]
  205. Wlodarski, M.W.; Collin, M.; Horwitz, M.S. GATA2 deficiency and related myeloid neoplasms. Semin. Hematol. 2017, 54, 81–86. [Google Scholar] [CrossRef] [PubMed]
  206. Spinner, M.A.; Sanchez, L.A.; Hsu, A.P.; Shaw, P.A.; Zerbe, C.S.; Calvo, K.R.; Arthur, D.C.; Gu, W.; Gould, C.M.; Brewer, C.C.; et al. GATA2 deficiency: A protean disorder of hematopoiesis, lymphatics, and immunity. Blood 2014, 123, 809–821. [Google Scholar] [CrossRef] [PubMed]
  207. Hsu, A.P.; McReynolds, L.J.; Holland, S.M. GATA2 deficiency. Curr. Opin. Allergy Clin. Immunol. 2015, 15, 104–109. [Google Scholar] [CrossRef] [PubMed]
  208. Shimamura, A. Aplastic anemia and clonal evolution: Germ line and somatic genetics. Hematol. Am. Soc. Hematol. Educ. Program 2016, 2016, 74–82. [Google Scholar] [CrossRef] [PubMed]
  209. Wlodarski, M.W.; Hirabayashi, S.; Pastor, V.; Stary, J.; Hasle, H.; Masetti, R.; Dworzak, M.; Schmugge, M.; van den Heuvel-Eibrink, M.; Ussowicz, M.; et al. Prevalence, clinical characteristics, and prognosis of GATA2-related myelodysplastic syndromes in children and adolescents. Blood 2016, 127, 1387–1397. [Google Scholar] [CrossRef] [PubMed]
  210. Park, M. Overview of inherited bone marrow failure syndromes. Blood Res. 2022, 57, 49–54. [Google Scholar] [CrossRef] [PubMed]
  211. Li, J.; Bledsoe, J.R. Inherited bone marrow failure syndromes and germline predisposition to myeloid neoplasia: A practical approach for the pathologist. Semin. Diagn. Pathol. 2023, 40, 429–442. [Google Scholar] [CrossRef] [PubMed]
  212. Deng, J.; McReynolds, L.J. Inherited bone marrow failure syndromes: A review of current practices and potential future research directions. Curr. Opin. Pediatr. 2023, 35, 75–83. [Google Scholar] [CrossRef]
  213. Dokal, I.; Vulliamy, T. Inherited bone marrow failure syndromes. Haematologica 2010, 95, 1236–1240. [Google Scholar] [CrossRef]
  214. Bhandari, J.; Thada, P.K.; Puckett, Y. Fanconi Anemia. In StatPearls; StatPearls: Treasure Island, FL, USA, 2023. [Google Scholar]
  215. Dufour, C.; Pierri, F. Modern management of Fanconi anemia. Hematol. Am. Soc. Hematol. Educ. Program 2022, 2022, 649–657. [Google Scholar] [CrossRef]
  216. Thakur, B.; Hiwale, K.M. Fanconi Anemia: A Rare Genetic Disorder. Cureus 2023, 15, e38899. [Google Scholar] [CrossRef] [PubMed]
  217. Da Costa, L.; Leblanc, T.; Mohandas, N. Diamond-Blackfan anemia. Blood 2020, 136, 1262–1273. [Google Scholar] [CrossRef] [PubMed]
  218. Da Costa, L.M.; Marie, I.; Leblanc, T.M. Diamond-Blackfan anemia. Hematol. Am. Soc. Hematol. Educ. Program 2021, 2021, 353–360. [Google Scholar] [CrossRef] [PubMed]
  219. Gadhiya, K.; Wills, C. Diamond Blackfan Anemia. In StatPearls; StatPearls: Treasure Island, FL, USA, 2023. [Google Scholar]
  220. AlSabbagh, M.M. Dyskeratosis congenita: A literature review. J. Dtsch. Dermatol. Ges. 2020, 18, 943–967. [Google Scholar] [CrossRef] [PubMed]
  221. Gitto, L.; Stoppacher, R.; Richardson, T.E.; Serinelli, S. Dyskeratosis congenita. Autops. Case Rep. 2020, 10, e2020203. [Google Scholar] [CrossRef] [PubMed]
  222. Garofola, C.; Nassereddin, A.; Gross, G.P. Dyskeratosis Congenita. In StatPearls; StatPearls: Treasure Island, FL, USA, 2023. [Google Scholar]
  223. Nelson, N.; Feurstein, S.; Niaz, A.; Truong, J.; Holien, J.K.; Lucas, S.; Fairfax, K.; Dickinson, J.; Bryan, T.M. Functional genomics for curation of variants in telomere biology disorder associated genes: A systematic review. Genet. Med. 2023, 25, 100354. [Google Scholar] [CrossRef]
  224. Cordell, V.; Osoba, L. Pregnancy in a patient with Schwachman-Diamond syndrome. BMJ Case Rep. 2015, 2015, bcr2015209644. [Google Scholar] [CrossRef]
  225. Woodward, E.R.; Meyer, S. Fanconi Anaemia, Childhood Cancer and the BRCA Genes. Genes 2021, 12, 1520. [Google Scholar] [CrossRef]
  226. Alter, B.P.; Giri, N.; Savage, S.A.; Peters, J.A.; Loud, J.T.; Leathwood, L.; Carr, A.G.; Greene, M.H.; Rosenberg, P.S. Malignancies and survival patterns in the National Cancer Institute inherited bone marrow failure syndromes cohort study. Br. J. Haematol. 2010, 150, 179–188. [Google Scholar] [CrossRef]
  227. van Dooijeweert, B.; Kia, S.K.; Dahl, N.; Fenneteau, O.; Leguit, R.; Nieuwenhuis, E.; van Solinge, W.; van Wijk, R.; Da Costa, L.; Bartels, M. GATA-1 Defects in Diamond-Blackfan Anemia: Phenotypic Characterization Points to a Specific Subset of Disease. Genes 2022, 13, 447. [Google Scholar] [CrossRef]
  228. Mello, F.V.; Land, M.G.P.; Costa, E.S.; Teodosio, C.; Sanchez, M.L.; Barcena, P.; Peres, R.T.; Pedreira, C.E.; Alves, L.R.; Orfao, A. Maturation-associated gene expression profiles during normal human bone marrow erythropoiesis. Cell Death Discov. 2019, 5, 69. [Google Scholar] [CrossRef] [PubMed]
  229. Savage, S.A.; Dufour, C. Classical inherited bone marrow failure syndromes with high risk for myelodysplastic syndrome and acute myelogenous leukemia. Semin. Hematol. 2017, 54, 105–114. [Google Scholar] [CrossRef] [PubMed]
  230. Feurstein, S.; Adegunsoye, A.; Mojsilovic, D.; Vij, R.; West DePersia, A.H.; Rajagopal, P.S.; Osman, A.; Collins, R.H.; Kim, R.H.; Gore, S.D.; et al. Telomere biology disorder prevalence and phenotypes in adults with familial hematologic and/or pulmonary presentations. Blood Adv. 2020, 4, 4873–4886. [Google Scholar] [CrossRef] [PubMed]
  231. Victorelli, S.; Passos, J.F. Telomeres and Cell Senescence—Size Matters Not. eBioMedicine 2017, 21, 14–20. [Google Scholar] [CrossRef] [PubMed]
  232. Myers, K.C.; Davies, S.M.; Shimamura, A. Clinical and molecular pathophysiology of Shwachman-Diamond syndrome: An update. Hematol. Oncol. Clin. N. Am. 2013, 27, 117–128. [Google Scholar] [CrossRef] [PubMed]
  233. Tan, S.; Kermasson, L.; Hoslin, A.; Jaako, P.; Faille, A.; Acevedo-Arozena, A.; Lengline, E.; Ranta, D.; Poiree, M.; Fenneteau, O.; et al. EFL1 mutations impair eIF6 release to cause Shwachman-Diamond syndrome. Blood 2019, 134, 277–290. [Google Scholar] [CrossRef]
  234. Godley, L.A. DNAJC21: The new kid on the SDS block. Blood 2017, 129, 1413–1414. [Google Scholar] [CrossRef]
  235. Federico, S.; Brennan, R.; Dyer, M.A. Childhood cancer and developmental biology a crucial partnership. Curr. Top. Dev. Biol. 2011, 94, 1–13. [Google Scholar] [CrossRef]
  236. Felix, C.A.; Lange, B.J. Leukemia in infants. Oncologist 1999, 4, 225–240. [Google Scholar] [CrossRef]
  237. Montes, R.; Ayllon, V.; Gutierrez-Aranda, I.; Prat, I.; Hernandez-Lamas, M.C.; Ponce, L.; Bresolin, S.; Te Kronnie, G.; Greaves, M.; Bueno, C.; et al. Enforced expression of MLL-AF4 fusion in cord blood CD34+ cells enhances the hematopoietic repopulating cell function and clonogenic potential but is not sufficient to initiate leukemia. Blood 2011, 117, 4746–4758. [Google Scholar] [CrossRef]
  238. Bursen, A.; Schwabe, K.; Ruster, B.; Henschler, R.; Ruthardt, M.; Dingermann, T.; Marschalek, R. The AF4.MLL fusion protein is capable of inducing ALL in mice without requirement of MLL.AF4. Blood 2010, 115, 3570–3579. [Google Scholar] [CrossRef] [PubMed]
  239. Bueno, C.; Ayllon, V.; Montes, R.; Navarro-Montero, O.; Ramos-Mejia, V.; Real, P.J.; Romero-Moya, D.; Arauzo-Bravo, M.J.; Menendez, P. FLT3 activation cooperates with MLL-AF4 fusion protein to abrogate the hematopoietic specification of human ESCs. Blood 2013, 121, 3867–3878. [Google Scholar] [CrossRef] [PubMed]
  240. Calo, E.; Wysocka, J. Modification of enhancer chromatin: What, how, and why? Mol. Cell 2013, 49, 825–837. [Google Scholar] [CrossRef] [PubMed]
  241. Valekunja, U.K.; Edgar, R.S.; Oklejewicz, M.; van der Horst, G.T.; O’Neill, J.S.; Tamanini, F.; Turner, D.J.; Reddy, A.B. Histone methyltransferase MLL3 contributes to genome-scale circadian transcription. Proc. Natl. Acad. Sci. USA 2013, 110, 1554–1559. [Google Scholar] [CrossRef] [PubMed]
  242. Kandoth, C.; McLellan, M.D.; Vandin, F.; Ye, K.; Niu, B.; Lu, C.; Xie, M.; Zhang, Q.; McMichael, J.F.; Wyczalkowski, M.A.; et al. Mutational landscape and significance across 12 major cancer types. Nature 2013, 502, 333–339. [Google Scholar] [CrossRef] [PubMed]
  243. Chen, C.; Liu, Y.; Rappaport, A.R.; Kitzing, T.; Schultz, N.; Zhao, Z.; Shroff, A.S.; Dickins, R.A.; Vakoc, C.R.; Bradner, J.E.; et al. MLL3 is a haploinsufficient 7q tumor suppressor in acute myeloid leukemia. Cancer Cell 2014, 25, 652–665. [Google Scholar] [CrossRef] [PubMed]
  244. Lee, J.; Kim, D.H.; Lee, S.; Yang, Q.H.; Lee, D.K.; Lee, S.K.; Roeder, R.G.; Lee, J.W. A tumor suppressive coactivator complex of p53 containing ASC-2 and histone H3-lysine-4 methyltransferase MLL3 or its paralogue MLL4. Proc. Natl. Acad. Sci. USA 2009, 106, 8513–8518. [Google Scholar] [CrossRef]
  245. Eguchi, M.; Eguchi-Ishimae, M.; Greaves, M. The role of the MLL gene in infant leukemia. Int. J. Hematol. 2003, 78, 390–401. [Google Scholar] [CrossRef]
  246. Valentine, M.C.; Linabery, A.M.; Chasnoff, S.; Hughes, A.E.; Mallaney, C.; Sanchez, N.; Giacalone, J.; Heerema, N.A.; Hilden, J.M.; Spector, L.G.; et al. Excess congenital non-synonymous variation in leukemia-associated genes in MLL-infant leukemia: A Children’s Oncology Group report. Leukemia 2014, 28, 1235–1241. [Google Scholar] [CrossRef]
  247. Arcipowski, K.M.; Bulic, M.; Gurbuxani, S.; Licht, J.D. Loss of Mll3 Catalytic Function Promotes Aberrant Myelopoiesis. PLoS ONE 2016, 11, e0162515. [Google Scholar] [CrossRef]
  248. Tawana, K.; Drazer, M.W.; Churpek, J.E. Universal genetic testing for inherited susceptibility in children and adults with myelodysplastic syndrome and acute myeloid leukemia: Are we there yet? Leukemia 2018, 32, 1482–1492. [Google Scholar] [CrossRef] [PubMed]
  249. Roloff, G.W.; Griffiths, E.A. When to obtain genomic data in acute myeloid leukemia (AML) and which mutations matter. Blood Adv. 2018, 2, 3070–3080. [Google Scholar] [CrossRef]
  250. Padella, A.; Ghelli Luserna Di Rora, A.; Marconi, G.; Ghetti, M.; Martinelli, G.; Simonetti, G. Targeting PARP proteins in acute leukemia: DNA damage response inhibition and therapeutic strategies. J. Hematol. Oncol. 2022, 15, 10. [Google Scholar] [CrossRef] [PubMed]
  251. Roloff, G.W.; Shaw, R.; O’Connor, T.E.; Hathaway, F.; Drazer, M.W. Stagnation in quality of next-generation sequencing assays for the diagnosis of hereditary hematopoietic malignancies. J. Genet. Couns. 2023, 32, 744–749. [Google Scholar] [CrossRef] [PubMed]
  252. Godley, L.A.; Shimamura, A. Genetic predisposition to hematologic malignancies: Management and surveillance. Blood 2017, 130, 424–432. [Google Scholar] [CrossRef]
  253. Carraway, H.E.; LaFramboise, T. Myeloid neoplasms with germline predisposition: Practical considerations and complications in the search for new susceptibility loci. Best Pract. Res. Clin. Haematol. 2020, 33, 101191. [Google Scholar] [CrossRef] [PubMed]
  254. Mujahed, H.; Jansson, M.; Bengtzen, S.; Lehamnn, S. Bone marrow stroma cells derived from mononuclear cells at diagnosis as a source of germline control DNA for determination of somatic mutations in acute myeloid leukemia. Blood Cancer J. 2017, 7, e616. [Google Scholar] [CrossRef] [PubMed]
  255. Kraft, I.L.; Godley, L.A. Identifying potential germline variants from sequencing hematopoietic malignancies. Blood 2020, 136, 2498–2506. [Google Scholar] [CrossRef]
  256. Saygin, C.; Roloff, G.; Hahn, C.N.; Chhetri, R.; Gill, S.; Elmariah, H.; Talati, C.; Nunley, E.; Gao, G.; Kim, A.; et al. Allogeneic hematopoietic stem cell transplant outcomes in adults with inherited myeloid malignancies. Blood Adv. 2023, 7, 549–554. [Google Scholar] [CrossRef]
  257. Williams, L.; Doucette, K.; Karp, J.E.; Lai, C. Genetics of donor cell leukemia in acute myelogenous leukemia and myelodysplastic syndrome. Bone Marrow Transplant. 2021, 56, 1535–1549. [Google Scholar] [CrossRef]
  258. Gibson, C.J.; Kim, H.T.; Zhao, L.; Murdock, H.M.; Hambley, B.; Ogata, A.; Madero-Marroquin, R.; Wang, S.; Green, L.; Fleharty, M.; et al. Donor Clonal Hematopoiesis and Recipient Outcomes after Transplantation. J. Clin. Oncol. 2022, 40, 189–201. [Google Scholar] [CrossRef] [PubMed]
  259. Pollyea, D.A.; Altman, J.K.; Assi, R.; Bixby, D.; Fathi, A.T.; Foran, J.M.; Gojo, I.; Hall, A.C.; Jonas, B.A.; Kishtagari, A.; et al. Acute Myeloid Leukemia, Version 3.2023, NCCN Clinical Practice Guidelines in Oncology. J. Natl. Compr. Cancer Netw. 2023, 21, 503–513. [Google Scholar] [CrossRef] [PubMed]
  260. de Haas, V.; Ismaila, N.; Advani, A.; Arber, D.A.; Dabney, R.S.; Patel-Donelly, D.; Kitlas, E.; Pieters, R.; Pui, C.H.; Sweet, K.; et al. Initial Diagnostic Work-Up of Acute Leukemia: ASCO Clinical Practice Guideline Endorsement of the College of American Pathologists and American Society of Hematology Guideline. J. Clin. Oncol. 2019, 37, 239–253. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Involvement of DDX41 variants in myeloid leukemogenesis. (A) Myeloid neoplasms arising from DDX41 variants: Hematopoietic cells carrying a heterozygous germline DDX41 variant (depicted as cells with blue nuclei) undergo the development of myeloid neoplasms following the acquisition of a somatic variant in the initially wild-type DDX41 after a prolonged latent period (illustrated as cells with light purple nuclei). The proportion of tumor cells tends to be low, and these cells may disrupt normal hematopoiesis, which is sustained by cells with only a germline variant. (B) Effects of R-loop accumulation on cellular function: R-loops form when transcribed RNA hybridizes with template DNA. The inappropriate accumulation of R-loops leads to DNA replication stress, impacting cellular function.
Figure 1. Involvement of DDX41 variants in myeloid leukemogenesis. (A) Myeloid neoplasms arising from DDX41 variants: Hematopoietic cells carrying a heterozygous germline DDX41 variant (depicted as cells with blue nuclei) undergo the development of myeloid neoplasms following the acquisition of a somatic variant in the initially wild-type DDX41 after a prolonged latent period (illustrated as cells with light purple nuclei). The proportion of tumor cells tends to be low, and these cells may disrupt normal hematopoiesis, which is sustained by cells with only a germline variant. (B) Effects of R-loop accumulation on cellular function: R-loops form when transcribed RNA hybridizes with template DNA. The inappropriate accumulation of R-loops leads to DNA replication stress, impacting cellular function.
Ijms 25 00652 g001
Figure 2. Role of p53 variants in cancer. p53 variants produce drug resistance, dominant negative effects on wild-type p53, proteasome repression, and LOF of wild-type p53. In cases of GOF, it promotes various cellular responses such as carcinogenesis, cancer cell proliferation, invasion, metastasis, tumor microenvironment establishment, genomic instability, and metabolic reprogramming.
Figure 2. Role of p53 variants in cancer. p53 variants produce drug resistance, dominant negative effects on wild-type p53, proteasome repression, and LOF of wild-type p53. In cases of GOF, it promotes various cellular responses such as carcinogenesis, cancer cell proliferation, invasion, metastasis, tumor microenvironment establishment, genomic instability, and metabolic reprogramming.
Ijms 25 00652 g002
Figure 3. Role of SAMD9 and SAMD9L in HSPC function. The SAMD9 and SAMD9L genes regulate proteins involved in the cell cycle, DNA damage repair, and ribosome regulation. Mutant SAMD9 and SAMD9L proteins significantly enhance these functions, which cause decreased hematopoietic potential and apoptosis in the bone marrow, promoting monosomy 7/del 7 HSPC production. Hematopoietic stem and progenitor cell (HSPC), myelodysplastic syndrome (MDS), and mutant type (MT).
Figure 3. Role of SAMD9 and SAMD9L in HSPC function. The SAMD9 and SAMD9L genes regulate proteins involved in the cell cycle, DNA damage repair, and ribosome regulation. Mutant SAMD9 and SAMD9L proteins significantly enhance these functions, which cause decreased hematopoietic potential and apoptosis in the bone marrow, promoting monosomy 7/del 7 HSPC production. Hematopoietic stem and progenitor cell (HSPC), myelodysplastic syndrome (MDS), and mutant type (MT).
Ijms 25 00652 g003
Table 1. Clinical characteristics, genetics, and prevalence of HHMS. Only the major genes discussed in this review are included in this table.
Table 1. Clinical characteristics, genetics, and prevalence of HHMS. Only the major genes discussed in this review are included in this table.
GeneChromosome
Location
Disorder NamePenetrance and
Lifetime Risk of HM
PrevalenceAge of MN Onset, YearsMalignancy TypesOther ManifestationsCitations
DDX415q35.3Familial MDS/AML
with mutated DDX41
incomplete penetranceUp to 13% of myeloid
neoplasms have a
genetic background,
of which DDX41
variants account for
about 80% of cases.
Median age is 65 years, ranging from 44 to 88
years, which notably
overlaps with the average
age of sporadic myeloid
malignancies.
MDS, AML, t-MN,
solid tumors,
especially colon and
prostate cancer and
melanoma, but not
yet definitively linked
cytopenia,
macrocytosis,
autoimmune diseases
[13,14,18,19,20,21,22,23]
TP5317p13.1Li-Fraumeni
syndrome (LFS)
lifetime risk of HM
is about 6%
LFS affects all
ethnicities and
has an estimated
incidence of
1:5000.
Nearly 100% of individuals
develop cancer by the age
of 70, with the median age
of first cancer at 20 to 30
years.
MDS, AML, ALL,
t-MN, lymphoma, MM,
osteosarcoma, breast
cancer, brain tumors,
soft tissue sarcoma,
adrenocortical
carcinoma and
other solid tumors
none[7,24,25,26,27,28,29]
CEBPA19q13.1Familial AML with
mutated CEBPA
>80% lifetime risk
of AML
<20 families
reported
Median age is 24.5 years,
ranging from 2 to 50 years.
AMLnone[13,30,31,32,33,34]
RUNX121q22.12Familial platelet
disorder with
propensity to
myeloid malignancy
unknown>250 families
reported
Median age is 33 years,
ranging from 6 to 76 years.
MDS, AML, ALL,
other lymphoid
malignancies
thrombocytopenia,
platelet dysfunction,
atopic and autoimmune
disorders
[13,35,36,37,38,39,40,41]
ANKRD2610p12.1Thrombocytopenia 2penetrance for
thrombocytopenia is
near complete, lifetime risk of HM is about 8%
UnknownMedian age is over 30 years,
ranging from 20s to 70s
years.
MDS, AML, CML,
MPN, ALL, CLL, MM
thrombocytopenia,
leukocytosis,
erythrocytosis,
mild bleeding tendency
[14,42,43,44]
ETV612p13.2Thrombocytopenia 5penetrance for
thrombocytopenia is
near complete
ALL is more
frequent, especially
in B-ALL (0.8% of
unselected childhood
B-cell ALL).
The ratio of lymphoid
versus myeloid
malignancies is
roughly 2:1.
Age ranges from 8 to 82 years
and seem to occur at
a younger age than usual
but is not yet defined.
ALL, MDS, AML,
CMML, MM,
GI cancers
thrombocytopenia,
macrocytosis,
platelet dysfunction
[13,15,45,46,47,48,49,50]
SAMD97q21.2MIRAGE Syndromeunknown8–17% of childhood
onset MDS
>110 individuals
reported
Pediatric age,
not yet defined.
MDS, AML, CMMLbone marrow failure,
cytopenia, infections,
growth restriction,
adrenal hypoplasia,
enteropathy, genital
abnormalities
[13,51,52,53,54,55]
SAMD9L7q21.2Ataxia, Pancytopenia
Syndrome
systemic
autoinflammatory
disease, bone marrow
failure, ataxia
GATA23q21.3GATA2 deficiency
syndrome
incomplete penetrance>480 individuals
reported, with 240 of
these confirmed to be
familial and 24 de
novo
Median age is 17 years,
ranging from 0 to 78 years.
MDS, AML, CMML,
ALL
immunodeficiency,
bone marrow failure, monocytopenia, lymphopenia, neutropenia, other cytopenia, infections, lymphedema,
congenital deafness, pulmonary alveolar proteinosis, venous and arterial thrombosis
[13,37,56,57,58]
ALL, acute lymphoblastic leukemia; CML, chronic myeloid leukemia; CMML, chronic myelomonocytic leukemia; CLL, chronic lymphocytic leukemia; t-MN, therapy-related myeloid neoplasms; MM, multiple myeloma; MPN, myeloproliferative neoplasm; HM, hematological malignancies.
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

Arai, H.; Matsui, H.; Chi, S.; Utsu, Y.; Masuda, S.; Aotsuka, N.; Minami, Y. Germline Variants and Characteristic Features of Hereditary Hematological Malignancy Syndrome. Int. J. Mol. Sci. 2024, 25, 652. https://doi.org/10.3390/ijms25010652

AMA Style

Arai H, Matsui H, Chi S, Utsu Y, Masuda S, Aotsuka N, Minami Y. Germline Variants and Characteristic Features of Hereditary Hematological Malignancy Syndrome. International Journal of Molecular Sciences. 2024; 25(1):652. https://doi.org/10.3390/ijms25010652

Chicago/Turabian Style

Arai, Hironori, Hirotaka Matsui, SungGi Chi, Yoshikazu Utsu, Shinichi Masuda, Nobuyuki Aotsuka, and Yosuke Minami. 2024. "Germline Variants and Characteristic Features of Hereditary Hematological Malignancy Syndrome" International Journal of Molecular Sciences 25, no. 1: 652. https://doi.org/10.3390/ijms25010652

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