Next Article in Journal
Effect of FKBP12-Derived Intracellular Peptides on Rapamycin-Induced FKBP–FRB Interaction and Autophagy
Next Article in Special Issue
Mediating EGFR-TKI Resistance by VEGF/VEGFR Autocrine Pathway in Non-Small Cell Lung Cancer
Previous Article in Journal
Physiological Function during Exercise and Environmental Stress in Humans—An Integrative View of Body Systems and Homeostasis
Previous Article in Special Issue
Role and Involvement of TENM4 and miR-708 in Breast Cancer Development and Therapy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Inosine Triphosphate Pyrophosphatase (ITPase): Functions, Mutations, Polymorphisms and Its Impact on Cancer Therapies

by
Mazin A. Zamzami
1,2
1
Department of Biochemistry, Faculty of Science, King Abdulaziz University, Jeddah 21589, Saudi Arabia
2
Centre of Artificial Intelligence in Precision Medicines, King Abdulaziz University, Jeddah 21589, Saudi Arabia
Submission received: 16 December 2021 / Revised: 18 January 2022 / Accepted: 19 January 2022 / Published: 24 January 2022
(This article belongs to the Special Issue Emerging Targets and Therapeutic Strategies in Cancer)

Abstract

:
Inosine triphosphate pyrophosphatase (ITPase) is an enzyme encoded by the ITPA gene and functions to prevent the incorporation of noncanonical purine nucleotides into DNA and RNA. Specifically, the ITPase catalyzed the hydrolysis of (deoxy) nucleoside triphosphates ((d) NTPs) into the corresponding nucleoside monophosphate with the concomitant release of pyrophosphate. Recently, thiopurine drug metabolites such as azathioprine have been included in the lists of ITPase substrates. Interestingly, inosine or xanthosine triphosphate (ITP/XTP) and their deoxy analogs, deoxy inosine or xanthosine triphosphate (dITP/dXTP), are products of important biological reactions such as deamination that take place within the cellular compartments. However, the incorporation of ITP/XTP, dITP/dXTP, or the genetic deficiency or polymorphism of the ITPA gene have been implicated in many human diseases, including infantile epileptic encephalopathy, early onset of tuberculosis, and the responsiveness of patients to cancer therapy. This review provides an up-to-date report on the ITPase enzyme, including information regarding its discovery, analysis, and cellular localization, its implication in human diseases including cancer, and its therapeutic potential, amongst others.

1. Introduction

Inosine triphosphate pyrophosphatase (ITPase) was first discovered in human erythrocytes in 1964 by Liakopoulou and Alivisatos. The tasks of ITPase were different from those of ATPase, competitively inhibited by adenine derivatives and dependent on the presence of magnesium (Mg2+). By using partially isolated ITPase, it was shown in the year 1970 by Vanderheide that the enzyme delivers PPi from human erythrocytes [1]. Almost a decade later, it was then distinguished in terms of enzyme kinetics for ITP, isolated 2-3000-fold from human erythrocytes, and studied to understand its work against ATP (Figure 1), GTP, CTP, and UTP. Until the beginning of the century, two major findings came in human ITPA. Firstly, in the very same year more attention was paid to the subcellular localization and tissue distribution of ITPA. It was found that bone marrow fibroblasts had significantly much higher activity in comparison with general erythrocytes. Secondly, the results from Vanderheiden demonstrate that ITPA was highly particular for ITP and dITP and also found that XTP was a substrate [2]. In the year 2001, several studies had been carried out to understand the expression of ITPA in E. coli. With the help of recombinant technology and northern blots, it was broadly studied in 24 tissues and notably enlisted in the pancreas, liver, heart, testis/ovary, thyroid gland, adrenal gland, etc. Purified recombinant ITPA created the accelerator and set the stage for future developments, as well as the identification of further substrates, such as the triphosphates of agent base analogues 6-hydroxyaminopurine (HAP) and 2-amino-6-hydroxyaminopurine [3].

2. ITPA: A Crucial Metabolic Enzyme

Inosine triphosphate pyrophosphatase is a ubiquitous protective key enzyme which regulates the contamination of cells at non-canonical levels. It does not allow the accumulation of nucleotides. The deamination of purine bases gives rise to inosine and xanthine nucleotides, which break down through ITP and PPi (Figure 2). Xanthine nucleotide (XTP) is also a substrate, but not much activity has been seen towards other nucleoside triphosphates, and in IDP or IMP it is completely absent (Figure 3 and Figure 4) [4]. ITPase helps in preventing the storage of ITP and lowers the chance of inosine nucleotides merging into nucleic acids. It was isolated, purified and characterized from human erythrocytes. With the help of a two-step colorimetric assay, the enzymatic properties of ITPase in human erythrocytes have been studied. It was found that in the cytosol portion the ATP and GTP deamination activities were not much enhanced [5]. The activity of ITPAs has been studied in different tissues such as bone marrow, skin fibroblasts, lymphoid lines, amniotic fluid, and erythrocytes. Average ITPase activity ranges between 4.9 and 294 units/mg protein. This enzyme has a broad range-specific activity in erythrocytes, from 15 to 722 pmol/h/g Hb [6]. Tested samples of leukocytes showed the lowest ITPase activity. Therefore, a lower presence or the absence of ITPase activity is a discrete characteristic which may be seen in erythrocytes and other types of cells.
Many studies have been carried out to measure the Km value for ITPA, and it was found that in every case, the ideal pH is alkaline and the need for divalent ion such as Mg2+ or Mn2+ is seen. A study of different samples of 6000 individuals revealed that seven samples showed that the deficit in ITPase activity was responsible for the presence of a high level of ITP. The short arm of chromosome 20 represents the gene that codes for ITPase, ITPA [7]. However, no genes have been cloned and characterized in mammalian erythrocytes; only a few rat tissues and the liver of rabbit. Structure-based identification with relevant biochemical analysis was revealed from Methanococcus jannaschii in a novel bacterial nucleoside triphosphate pyrophosphate, Mj0226. A protein found in yeast, named Ham 1p, has been shown to be the result of a gene controlling the response of yeast strains to 6-N-hydroxylaminopurine (HAP). With about 30% sequence identity, this yeast protein is analogous to j0226 protein [8].
The ITPase is an α/β homodimeric protein that contains 194 amino acid residues, which form a 45 kDa dimer [5]. The two globular lobes of the enzyme are supported by a central elongated mixed β sheet. The binding of ITP takes place between the lobes, which are adjacent to the dimer interface. This homodimer contains a catalytic site which is found towards the periphery and a specificity pocket which is deeply buried in the dimerization of each monomer lobe. The particular mechanism has not yet been explained. However, it is believed that Mg2+ is needed for the catalytic activity, which takes place via acid–base chemistry. It has been seen that ITPase works best in the presence of reducing agents such as DTT in an alkaline pH. ITPase does not differentiate between ribose and deoxyribose sugar and breaks down phosphoanhydride attachment in non-canonical NTPs and dNTPs with the same interest [5,7].
The protein substrate complex has been elucidated through X-ray crystal structure. It was found that there are some unfolded preserved remains that indicate the control of substrate particularity. Almost 40 years later, the exact mechanism of the ITPase in distinguishing between canonical and noncanonical (d)NTPS has still not been completely revealed. Competitive inhibition has been seen only in human ITPase by inosine 5′-diphosphate; it has not been seen in NMPs or any other trialled nucleoside. A study showed the strong inhibition of enzymes by ions such as Ca2+, Cd2+, and Co2+ [5,7]. The only nucleoside tested was inosine 5′-diphosphate. Competitive inhibition is shown in human ITPase at inosine 5′-diphosphate. However, positive cooperativity is seen only in the E. coli ITPase ortholog, RdgB [9]. To date, ITPase biochemistry has been overlooked. However, some advanced studies have been conducted regarding substrate selectivity and clinically relevant variants. ITPase plays a very important role in purine metabolism. However, research gaps exist in understanding the biochemistry and function of ITPase. There are several roles of ITPase which are still unknown, such as the role of ITP in cellular substrate inhibition. It effects the regulation of the concentration of ITP, IPM, or inosine inside the cell, and whether the enzyme is allosterically regulated or post-translationally modified [9].
Substrate inhibition has been studied in different sources, and it was found that when a protein source comes from the whole-genome extract of erythrocyte cells, it is not subjected to substrate inhibition [10]. However, in enzymology, ITPase is defined as a source of protein inhibition. The central idea of this study tells us that the substrate binding of ITPase activity may have a pivotal role in the development of therapeutics.

3. ITPA Mutations and Association with Clinical Disease

It is believed that mutation significantly affects the structure of a protein. It mostly causes a change in the length of 194 residues of amino acids, and may result in non-functional proteins. Patients with a homozygous mutation as a result of deletion, duplication or frameshift mutation in erythrocyte or fibroblast cells exhibit severe damage to or the complete failure of protein activity [11]. Out of seven clinically relevant ITPA variants, one is a duplication, the other a nonsense mutation and a deletion, and the remaining four are single-nucleotide substitutions. Nonsense, duplication and deletion mutations are thought to be against variants and are very infrequent. All three mutations were pointed out in a group of patients with infantile encephalopathy.
In mutants (c.359_366dupTCAGCACC), a duplication mutation of 8 base pairs results in a frameshift mutation at the 123 position of an amino acid, which was speculated to produce 225 amino acid polypeptides with changes in amino acid sequence. A deletion of 1874 base pairs was seen in a mutant (c.264-607_295 + 1267del), which completely spanned exon 5 and resulted in a frame transcript. A premature stop codon at the 151 position of the amino acid is introduced as a result of a nonsense mutation in c.452G > A, which results in the shortening of C-terminal protein, which may tremendously preserve the specificity pocket that contains SHR, a trademark sequence at positions from 176 to 178 [4,11].
Out of the seven nucleotide mutants listed, four are clinically relevant and the most studied variant is the c.94C > A (p.Pro32Thr). Two single-nucleotide mutant/variants affect protein structure in several ways, including reduced catalytic activity and decreased stability and expression of the full-length transcript [11]. This type of point mutation is responsible for mRNA splicing events at exon 2 and 3, which results in a non-functional protein. This event is reduced in wild-type cells, while this increases almost 3-fold in homozygous mutants. Other studies showed that the replacement of proline with threonine at position 32 is responsible for proteins having a reduced stability and lesser catalytic activity [4]. Furthermore, the ITPA c.94C > A sequence variant has been shown to be linked with the susceptibility to advanced drug reactions in azathioprine-treated patients. In addition, the ITPA c.94C > A allelic variant has been proposed as being responsible for destroying the exonic splicing silencing element in exon 2 in peripheral blood leukocytes patients; ultimately, this led to a change in the structure of the ITPase and contributed to its deficiency (the ITPA c.94C > A and c.124 + 21A > C (g.IVS2 + 21A > C) sequence variants led to misplacing of the ITPA gene) [12]. In adult hematological malignancy patients, the ITPA 94C > A variant has been reported to be linked to a substantial surge in the total heteroplasmic/homoplasmic mutations in the mitochondrial DNA, which implies that a decrease in the activity of the ITPase may likely give rise to changes in the mitochondrial ITPA and a possible association with mitochondrial DNA defects [13].
The c.532C > T (p.Arg178Cys) mutant is infrequent and was introduced in patients suffering from infantile encephalopathy. One of the residues that form the SHR signature sequence for ITPase is Arg-178, which is highly conserved in the ITPase. In vivo and in vitro site-directed mutagenesis studies have shown that this position is prime for ITPase activity. Additionally, with the help of many software programs, it is supposed to link with the nucleobase of incoming (d)ITP at two positions, which is thought to result in a non-functional protein. The other two nucleotide mutants do not affect the protein structure [14].
Additionally, it was found that ITPase deficiency is linked to the severity of organ dysfunction. For example, in families affected by ITPA mutations involving both alleles, these people expressed muscle heart failure, which was detrimental at an early age. They also possessed features that resembled Martsolf syndrome, a genetic disorder that affects the eye and the brain (ITPase deficiency leads to Martsolf-like syndrome with a lethal infantile dilated cardiomyopathy) [12]. The severity of the ITPA loss has been tested in mice. For example, ITPA knockout in a mice model leads to the accumulation of inosine nucleotides in both the nucleotide and RNA poll, resulting in the death of the mice at infancy. These mice showed a distinctive feature that resembled growth retardation and the disorganization of the cardiac myofiber (ITPase-deficient mice show growth retardation and die before weaning) [15].

4. ITPA Mutations and Therapeutic Implications

In the last few years, the pharmacogenetic importance of ITPA mutations has been reported. To date, studies suggest that a minimum of 30 percent polymorphism is identified in the ITPA gene. Of thirty studies, seven were demonstrated to be clinically suitable (Table 1).
According to these studies, the ITPA position influences the results of thiopurine therapy and hepatitis C treatment [11]. ITPA mutation is also associated with young-onset tuberculosis susceptibility and is responsible for early infantile encephalopathy. Detailed study of the physical expression of the genes associated with ITPA has shown a lot of advancement in clinical treatment [4].
In the year 2009, it was found that the Asian population retains an ITPA variation of about 14% to 19% which is the highest among the total 5% of global inhabitants that hold onto c.94C > A (p.Pro32Thr) ITPA variation [16]. Subsequently, many other kinds of polymorphisms have been discovered. Clinical mutations are very occasional; in fact, data suggest that the low ITPase activity is responsible for the ITPA polymorphism in almost 75% of the affected population [13]. Variation in the genotype may result in a downfall in various levels of ITPase activity. It is crucial to understand that the results prevail from erythrocytes and the ITPase activity measured is different for different cell types, even if it is from similar individuals. Various studies have been carried out to understand the ITPase activity and common types of polymorphism in erythrocytes c.94C > A (p.Pro32Thr) and c.124 + 21A > C (g.IVS2 + 21A > C). A polymorphism in Caucasian populations, the c.94C > A (p.Pro32Thr), shows a serious downfall in ITPase activity and the average activity for heterozygous individuals is ~25 %, whereas homozygous individuals maintain less than 1% of the wild-type levels [14,17]. Moreover, an IPTA genotyping study in a quarter of the Tunisian population showed that polymorphisms in that population had reduced metabolic functioning of the enzyme [18]. It was also found that the Azathioprine-induced neutropoenia in PR3+ microscopic polyangiitis was probably due to an IPTA 94C > A mutation [19].
The c.124 + 21A > C variant results in a small reduction in ITPase activity, and heterozygotes maintain about 60% of the wild-type levels while homozygous individuals maintain about 30 % of the wild-type levels [4]. The resulting c.94C > A (p.Pro32Thr) and c.124 + 21A > C compound heterozygosity demonstrate a major reduction in ITPase activity and allowed them to maintain 8% of the wild-type levels [16]. Previous studies have been conducted in erythrocytes in order to understand the common polymorphisms of ITPase activity, c.94C > A (p.Pro32Thr) and c.124 + 21A > C. The current spectrum of genotype status shows a decrease in ITPase activity [20].

4.1. Infantile Encephalopathy

Recently, Kevelam et al. found that one of the reasons behind early infantile encephalopathy is a recessive ITPA mutation. With the help of whole-exome sequencing (WES) and magnetic resonance imaging (MRI), researchers came to know of a distinctive pattern of MRI, which shows a recessive mutation of the ITPA gene [11]. This study was conducted on seven patients and six patients died before 2.5 years of age. Those patients faced seizures, developmental delay, and very serious progressive microcephaly from birth. Of the different profiling methods, the tests samples showed normal levels of purine and pyrimidine. However, it affected the heart function and resulted in a lower number of RBCs. Homozygous genetic mutation is one of the main reasons behind this, including missense mutations, frameshift mutations, nonsense mutations, and gene deletions. Missense mutations are responsible for the change in the primary structure of the amino acid. In such cases, it is observed that Arg-178 is replaced with cysteine [11,21]. Moreover, a study also showed that ITPase deficiency may lead to refractory epilepsy, microcephaly, and neurodevelopmental disease [22]. The effect of ITPase deficiency on neural epilepsy has been confirmed in knockout mice [23], which is also supported by another mice study that showed ITPase deficiency may lead to growth retardation [24].

4.2. Cancer Chemotherapy (Thiopurine Treatment)

A different group of studies showed that ITPA variation has a direct role in the increased toxicity of thiopurines. It is important to understand that not each and every ITPA polymorphism is related to thiopurine toxicity. Nowadays, thiopurines such as azathioprine or 6-mercaptopurine are extensively used to treat several diseases, such as inflammatory bowel diseases, ulcerative colitis, Crohn’s disease, cancer, and organ transplants (29,30). Patients with ITPA mutation have many mild to severe side effects, which include rashes, liver toxicity, inflammation of the pancreas, and aplastic anaemia. The presence of thiol-contacting NTPs is responsible for such toxic conditions. These life-threatening side effects sometimes result in either discontinuation or dosage alteration. In order to minimize drug adversity, scientists have suggested the inclusion of pre-screening of the drug for use in patients affected by ITPA polymorphism [11,25]. A study also showed that aberrant activity of ITPase caused the accumulation of non-canonical nucleotides that may induce DNA damage and mutagenesis (Figure 5) [26], which was also supported by another study that confirmed ITPA as having a key role in humans to protect DNA [27].
Patients suffering from Crohn’s disease show resistance to 6-mercaptopurine therapy, which is successfully replaced using 6-thioguanine [14]. The main reason behind thiopurine toxicity is still unknown. Some studies do not suggest any correlation between ITPA mutation and drug toxicity, and there is still no specific set of rules that have been identified between the ITPA variation and drug toxicity. However, potential factors could include race, sample size, or drug reaction. Instead of azathioprine or 6-mercaptopurine, it is encouraged to used 6-thioguanine. In 2014, Matimba et al. demonstrated an experimental method called the “three-tiered” strategy to reduce the failure related to the clinical use of this drug [28].
In addition, it was previously found that the combined treatment of thiopurine with anti-tumour necrosis factor (TNF) can minimize the risk of anti-drug antibody formation that may attenuate the anti-TNF agent’s response. The effectiveness of thiopurine is different according to racial differences. For instances, the effectiveness of thiopurine treatment was found to be significant in East Asian populations, including Korea, China and Japan [29]. The occurrence of leukopenia is lower in Caucasian populations than in Asian populations, and hair loss is exceptional in Caucasians and not abnormal in Japanese patients, while the standard dose of thiopurines in Europe (AZA 2–2.5 mg/kg/day) and in Japan (AZA 1–2 mg/kg/day) differs [15,30]. Additionally, a study also showed that silencing the ITPase gene led to the induction of apoptosis in folate single-wall-nanotube-treated SKBR3 cancer cells [31].
The genetic polymorphism of the enzymes affects the metabolism of thiopurine. A variation in individual thiopurine metabolism is responsible for some of the adverse reactions. Pharmacogenetic studies have been reported in leukaemia, organ transplantation and IBD and some pharmacogenetic predictors have been found and are being used in clinical practice [25,32].

4.3. Tuberculosis Treatment

Recently, it was shown that ITPA overexpression is implicated in juvenile tuberculosis. By using next-generation sequencing techniques, ITPA polymorphism was identified in multiple juvenile patients from different families suffering from tuberculosis, with the expression of g.19176G > A and c.94C > A (p.Pro32Thr) related to juvenile TB patients [4]. Among many variants of the ITPA gene, the g.19176G > variant located at the 3′-UTR demonstrates the highest bond and is supposed to enhance the expression of ITPA at the post-transcriptional level. A minor ‘A’ allele form of this variant demonstrates a higher level of expression by using a lymphoblastic expression profile (in silico) [4,33]. Eventually, it was speculated that the lower level of expression for the major ‘G’ allele may cause individuals with this allele to be more susceptible to TB infection [4]. This study reveals how the formulation of ITPA can play a defensive role in the onset of TB. The human thymus tissue is found to express the highest amount of ITPA and the results found after the detailed study of ITPA expression in the thymus gland show that it may have a very strong role in the development and outbreak of tuberculosis [4].

4.4. Hepatitis C Treatment

Around the globe, approximately 170 million patients are facing hepatitis C, and one of the strongest offshoots after treatment with interferon alpha and ribavirin is haemolytic anaemia [33]. Similarly, ITPase polymorphism (rs1127354) was found to be associated with haemoglobin level in treated Chinese cohorts of hepatitis C [34]. Various studies showed that the variant c.94C > A (p.Pro32Thr) is responsible for delaying the conditions responsible for anaemia. Data show that patients having variants c.94C > A (p.Pro32Thr) or c.124 + 21A > C not only develop the symptoms of anaemia, but also require a low reduction in riboflavin dose [35,36,37]. Nowadays, the ITPA activity is used as a probe to detect the development of anaemia in patients taking riboflavin treatment [38]. Based on the above finding, it is advisable to use the ITPase activity or ITPA locus in the treatment of hepatitis C using ribavirin [39].

4.5. Antiviral Treatment-Driven Anaemia

Apart from that mentioned above, ITPase deficiency was found to have a major impact on the patients treated with antivirals. In general, the patients with a moderate deficiency showed the most protection during the complete course of the antiviral therapy. Under these scenarios, the rate of haemoglobin decline was severe in patients with wild-type ITPase activity. This was found to be highly correlated with a greater deficiency in haemoglobin levels over the full course of therapy.
Overall, the abovementioned shows that ITPA mutations have a strong association with treatments for a variety of diseases, including cancer.

5. ITPA Variants/Mutants: A Disease Range

With the help of different studies, it is now very clear that ITPase plays a very important role in the purine metabolism pathway and ITPA variation is responsible for modulated ITPA activity/expression, which causes different diseases in different parts of the body [40,41]. Firstly, it was identified by Vanderheiden in patients with the abnormal accumulation of ITP in erythrocytes, and it was suggested that the patients were ITPase deficient [40,42]. However, with more research, in the year 2002, it was shown that there is no harm in the absence of ITPase activity, with zero percent activity from c.94C > A (p.Pro32Thr) homozygous individuals [11,33]. Additionally, a study also showed that ITPase gene variants prevent haemolytic anaemia in treated HCV patients [42]. Today it is understood that the role of this protein is different in all the different tissues of the body. This underscores that there is a very important relationship between the ITPA variation and drug-dosage system.

6. Final Remarks

In view of the ongoing evidence finding that the phenotypic series of ITPA defects ranges from very mild to severe deaths, it was also suggested that therapies will be developed to regulate ITPase activity. The ITPA status contributes to several severe diseases, from cardiomyopathy to neural defects and disorders of the immune system, as well as several other ITPA-related diseases [4,22,23]. The low-cost development of sequencing techniques, such as NGS and WES, will help to evaluate several other ITPase-related diseases.

7. Conclusions

There is a need for further research in order to understand the ITPA polymorphism. To date, only seven mutants/variants are known and this has posed a great challenge to clinicians to understand it completely. Several previous studies have indeed provided valuable inputs to increase our understanding of the mechanism involved in the differentiation of canonical and non-canonical purine metabolism and its involvement in disease. Advances in this field, taking the shape of continuous progress, have recently witnessed quantum leaps, but there is still much to study.

Funding

No funding was available.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We apologize to those researchers whose work was not included in this review due to space limitations. We would like to thank the Deanship of Scientific Research (DSR) at King Abdulaziz University (KAU), Jeddah, KSA for providing support in the collection of critical articles in the completion of this review.

Conflicts of Interest

There are no competing interests to be declared.

References

  1. Bierau, J.; Lindhout, M.; Bakker, J.A. Pharmacogenetic significance of inosine triphosphatase. Pharmacogenomics 2007, 8, 1221–1228. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Fraser, J.H.; Meyers, H.; Henderson, J.F.; Brox, L.W.; McCoy, E.E. Individual variation in inosine triphosphate accumulation in human erythrocytes. Clin. Biochem. 1975, 8, 353–364. [Google Scholar] [CrossRef]
  3. Vanderheiden, B.S. ITP pyrophosphohydrolase and idp phosphohydrolase in rat tissue. J. Cell. Physiol. 1975, 86, 167–175. [Google Scholar] [CrossRef] [PubMed]
  4. Burgis, N.E. A disease spectrum for ITPA variation: Advances in biochemical and clinical research. J. Biomed. Sci. 2016, 23, 73. [Google Scholar] [CrossRef] [Green Version]
  5. Holmes, S.L.; Turner, B.M.; Hirschhorn, K. Human inosine triphosphatase: Catalytic properties and population studies. Clin. Chim. Acta 1979, 97, 143–153. [Google Scholar] [CrossRef]
  6. Vanderheiden, B.S. Human erythrocyte “ITPase”: An ITP pyrophosphohydrolase. Biochim. Biophys. Acta-Gen. Subj. 1970, 215, 555–558. [Google Scholar] [CrossRef]
  7. Vanderheiden, B.S. Purification and properties of human erythrocyte inosine triphosphate pyrophosphohydrolase. J. Cell. Physiol. 1979, 98, 41–47. [Google Scholar] [CrossRef]
  8. Burgis, N.E.; Cunningham, R.P. Substrate Specificity of RdgB Protein, a Deoxyribonucleoside Triphosphate Pyrophosphohydrolase. J. Biol. Chem. 2007, 282, 3531–3538. [Google Scholar] [CrossRef] [Green Version]
  9. Porta, J.; Kolar, C.; Kozmin, S.G.; Pavlov, Y.I.; Borgstahl, G.E.O. Structure of the orthorhombic form of human inosine triphosphate pyrophosphatase. Acta Crystallogr. Sect. F Struct. Biol. Cryst. Commun. 2006, 62, 1076–1081. [Google Scholar] [CrossRef] [PubMed]
  10. Stenmark, P.; Kursula, P.; Flodin, S.; Gräslund, S.; Landry, R.; Nordlund, P.; Schüler, H. Crystal Structure of Human Inosine Triphosphatase: Substrate Binding and Implication of the Inosine Triphosphatase Deficiency Mutation P32t. J. Biol. Chem. 2007, 282, 3182–3187. [Google Scholar] [CrossRef] [Green Version]
  11. Kevelam, S.H.; Salvarinova, R.; Agrawal, S.; Visser, D.; Weiss, M.M.; Abbink, T.E.M.; Waisfisz, Q.; Bierau, J.; Honzík, T.; Salomons, G.S.; et al. Recessive ITPA mutations cause an early infantile encephalopathy. Ann. Neurol. 2015, 78, 649–658. [Google Scholar] [CrossRef] [PubMed]
  12. Arenas, M.; Duley, J.; Sumi, S.; Sanderson, J.; Marinaki, A. The ITPA c. 94C > A and g. IVS2 + 21A > C sequence variants contribute to missplicing of the ITPA gene. Biochim. Biophys. Acta-Mol. Basis Dis. 2007, 1772, 96–102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Herting, G.; Barber, K.; Zappala, M.R.; Cunningham, R.P.; Burgis, N.E. Quantitative in vitro and in vivo characterization of the human P32T mutant ITPase. Biochim. Biophys. Acta-Mol. Basis Dis. 2010, 1802, 269. [Google Scholar] [CrossRef]
  14. Shipkova, M.; Franz, J.; Abe, M.; Klett, C.; Wieland, E.; Andus, T. Association Between Adverse Effects Under Azathioprine Therapy and Inosine Triphosphate Pyrophosphatase Activity in Patients with Chronic Inflammatory Bowel Disease. Ther. Drug Monit. 2011, 33, 321–328. [Google Scholar] [CrossRef] [PubMed]
  15. Zamzami, M.A.; Duley, J.A.; Price, G.R.; Venter, D.J.; Yarham, J.W.; Taylor, R.W.; Catley, L.P.; Florin, T.H.; Marinaki, A.M.; Bowling, F. Inosine Triphosphate Pyrophosphohydrolase (ITPA) polymorphic sequence variants in adult hematological malignancy patients and possible association with mitochondrial DNA defects. J. Hematol. Oncol. 2013, 6, 24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Marsh, S.; King, C.R.; Ahluwalia, R.; McLeod, H.L. Distribution of ITPA P32T alleles in multiple world populations. J. Hum. Genet. 2004, 49, 579–581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Chadli, Z.; Kerkeni, E.; Hannachi, I.; Chouchene, S.; Ben Fredj, N.; Boughattas, N.A.; Aouam, K.; Chaabane, A. Distribution of Genetic Polymorphisms of Genes Implicated in Thiopurine Drugs Metabolism. Ther. Drug Monit. 2018, 40, 655–659. [Google Scholar] [CrossRef]
  18. Honda, K.; Kobayashi, A.; Niikura, T.; Hasegawa, T.; Saito, Z.; Ito, S.; Sasaki, T.; Komine, K.; Ishizuka, S.; Motoi, Y.; et al. Neutropenia related to an azathioprine metabolic disorder induced by an inosine triphosphate pyrophosphohydrolase (ITPA) gene mutation in a patient with PR3-ANCA-positive microscopic polyangiitis. Clin. Nephrol. 2018, 90, 363–369. [Google Scholar] [CrossRef]
  19. Rembeck, K.; Waldenström, J.; Hellstrand, K.; Nilsson, S.; Nyström, K.; Martner, A.; Lindh, M.; Norkrans, G.; Westin, J.; Pedersen, C.; et al. Variants of the inosine triphosphate pyrophosphatase gene are associated with reduced relapse risk following treatment for HCV genotype 2/3. Hepatology 2014, 59, 2131–2139. [Google Scholar] [CrossRef] [Green Version]
  20. Kaur, P.; Neethukrishna, K.; Kumble, A.; Girisha, K.M.; Shukla, A. Identification of a novel homozygous variant confirms ITPA as a developmental and epileptic encephalopathy gene. Am. J. Med. Genet. Part A 2019, 179, 857–861. [Google Scholar] [CrossRef]
  21. Scala, M.; Wortmann, S.B.; Kaya, N.; Stellingwerff, M.D.; Pistorio, A.; Glamuzina, E.; van Karnebeek, C.D.; Skrypnyk, C.; Iwanicka-Pronicka, K.; Piekutowska-Abramczuk, D.; et al. Clinico-radiological features, molecular spectrum, and identification of prognostic factors in developmental and epileptic encephalopathy due to inosine triphosphate pyrophosphatase (ITPase) deficiency. Hum. Mutat. 2022. [Google Scholar] [CrossRef]
  22. Koga, Y.; Tsuchimoto, D.; Hayashi, Y.; Abolhassani, N.; Yoneshima, Y.; Sakumi, K.; Nakanishi, H.; Toyokuni, S.; Nakabeppu, Y. Neural stem cell–specific ITPA deficiency causes neural depolarization and epilepsy. JCI Insight 2020, 5, e140229. [Google Scholar] [CrossRef] [PubMed]
  23. Behmanesh, M.; Sakumi, K.; Abolhassani, N.; Toyokuni, S.; Oka, S.; Ohnishi, Y.N.; Tsuchimoto, D.; Nakabeppu, Y. ITPase-deficient mice show growth retardation and die before weaning. Cell Death Differ. 2009, 16, 1315–1322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. von Ahsen, N.; Armstrong, V.W.; Behrens, C.; von Tirpitz, C.; Stallmach, A.; Herfarth, H.; Stein, J.; Bias, P.; Adler, G.; Shipkova, M.; et al. Association of Inosine Triphosphatase 94C > A and Thiopurine S-Methyltransferase Deficiency with Adverse Events and Study Drop-Outs under Azathioprine Therapy in a Prospective Crohn Disease Study. Clin. Chem. 2005, 51, 2282–2288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Menezes, M.R.; Waisertreiger, I.S.-R.; Lopez-Bertoni, H.; Luo, X.; Pavlov, Y.I. Pivotal Role of Inosine Triphosphate Pyrophosphatase in Maintaining Genome Stability and the Prevention of Apoptosis in Human Cells. PLoS ONE 2012, 7, e32313. [Google Scholar] [CrossRef] [Green Version]
  26. Waisertreiger, I.S.-R.; Menezes, M.R.; Randazzo, J.; Pavlov, Y.I. Elevated Levels of DNA Strand Breaks Induced by a Base Analog in the Human Cell Line with the P32T ITPA Variant. J. Nucleic Acids 2010, 2010, 872180. [Google Scholar] [CrossRef] [Green Version]
  27. Zelinkova, Z.; Derijks, L.J.; Stokkers, P.C.; Vogels, E.W.; van Kampen, A.H.; Curvers, W.L.; Cohn, D.; van Deventer, S.J.; Hommes, D.W. Inosine Triphosphate Pyrophosphatase and Thiopurine S-Methyltransferase Genotypes Relationship to Azathioprine-Induced Myelosuppression. Clin. Gastroenterol. Hepatol. 2006, 4, 44–49. [Google Scholar] [CrossRef]
  28. Matimba, A.; Li, F.; Livshits, A.; Cartwright, C.S.; Scully, S.; Fridley, B.L.; Jenkins, G.; Batzler, A.; Wang, L.; Weinshilboum, R.; et al. Thiopurine pharmacogenomics: Association of SNPs with clinical response and functional validation of candidate genes. Pharmacogenomics 2014, 15, 433–447. [Google Scholar] [CrossRef] [Green Version]
  29. Koren, G.; Ferrazini, G.; Sulh, H.; Langevin, A.M.; Kapelushnik, J.; Klein, J.; Giesbrecht, E.; Soldin, S.; Greenberg, M. Systemic Exposure to Mercaptopurine as a Prognostic Factor in Acute Lymphocytic Leukemia in Children. N. Engl. J. Med. 2010, 323, 17–21. [Google Scholar] [CrossRef]
  30. Charbgoo, F.; Behmanesh, M.; Nikkhah, M.; Kane, E.G. RNAi mediated gene silencing of ITPA using a targeted nanocarrier: Apoptosis induction in SKBR3 cancer cells. Clin. Exp. Pharmacol. Physiol. 2017, 44, 888–894. [Google Scholar] [CrossRef]
  31. Derijks, L.; Wong, D.R. Pharmacogenetics of Thiopurines in Inflammatory Bowel Disease. Curr. Pharm. Des. 2010, 16, 145–154. [Google Scholar] [CrossRef] [PubMed]
  32. Nakauchi, A.; Wong, J.H.; Mahasirimongkol, S.; Yanai, H.; Yuliwulandari, R.; Mabuchi, A.; Liu, X.; Mushiroda, T.; Wattanapokayakit, S.; Miyagawa, T.; et al. Identification of ITPA on chromosome 20 as a susceptibility gene for young-onset tuberculosis. Hum. Genome Var. 2016, 3, 15067. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Chi, X.; Wang, M.; Pan, Y.; Jiang, J.; Jiang, T.; Yan, H.; Wu, R.; Wang, X.; Gao, X.; Niu, J. Inosine triphosphate pyrophosphatase polymorphisms are predictors of anemia in Chinese patients with chronic hepatitis C during therapy with ribavirin and interferon. J. Gastroenterol. Hepatol. 2020, 35, 97–103. [Google Scholar] [CrossRef] [PubMed]
  34. Pineda-Tenor, D.; Garcia-Alvarez, M.; Jimenez-Sousa, M.A.; Vazquez-Moron, S.; Resino, S. Relationship between ITPA polymorphisms and hemolytic anemia in HCV-infected patients after ribavirin-based therapy: A meta-analysis. J. Transl. Med. 2015, 13, 320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Lavanchy, D. The global burden of hepatitis C. Liver Int. 2009, 29 (Suppl. 1), 74–81. [Google Scholar] [CrossRef]
  36. Chen, S.H.; Peng, C.Y.; Lai, H.C.; Su, W.P.; Lin, C.H.; Li, Y.F.; Chuang, P.H.; Chen, C.H. An Index to Predict Ribavirin-Induced Anemia in Asian Patients with Chronic Genotype 1 Hepatitis C. Zahedan J. Res. Med. Sci. 2015, 15, e27148. [Google Scholar] [CrossRef] [Green Version]
  37. Ampuero, J.; Romero-Gómez, M. Pharmacogenetics of ribavirin-induced anemia in hepatitis C. Pharmacogenomics 2016, 17, 1587–1594. [Google Scholar] [CrossRef]
  38. Thompson, A.J.; Fellay, J.; Patel, K.; Tillmann, H.L.; Naggie, S.; Ge, D.; Urban, T.J.; Shianna, K.V.; Muir, A.J.; Fried, M.W.; et al. Variants in the ITPA Gene Protect Against Ribavirin-Induced Hemolytic Anemia and Decrease the Need for Ribavirin Dose Reduction. Gastroenterology 2010, 139, 1181–1189. [Google Scholar] [CrossRef] [Green Version]
  39. Vanderheiden, B.S. Genetic studies of human erythrocyte inosine triphosphatase. Biochem. Genet. 1969, 3, 289–297. [Google Scholar] [CrossRef]
  40. Maeda, T.; Sumi, S.; Ueta, A.; Ohkubo, Y.; Ito, T.; Marinaki, A.M.; Kurono, Y.; Hasegawa, S.; Togari, H. Genetic basis of inosine triphosphate pyrophosphohydrolase deficiency in the Japanese population. Mol. Genet. Metab. 2005, 85, 271–279. [Google Scholar] [CrossRef] [PubMed]
  41. Vanderheiden, B.S. Inosine Triphosphate in Human Erythrocytes: A Genetic Trait. In International Society of Blood Transfusion; Karger Publishers: Basel, Switzerland, 1965; Volume 23, pp. 540–548. [Google Scholar] [CrossRef]
  42. Fellay, J.; Thompson, A.J.; Ge, D.; Gumbs, C.E.; Urban, T.J.; Shianna, K.V.; Little, L.D.; Qiu, P.; Bertelsen, A.H.; Watson, M.; et al. ITPA gene variants protect against anaemia in patients treated for chronic hepatitis C. Nature 2010, 464, 405–408. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structure of Inosine triphosphate (ITP).
Figure 1. Chemical structure of Inosine triphosphate (ITP).
Cells 11 00384 g001
Figure 2. Chemical formation of ITP molecule.
Figure 2. Chemical formation of ITP molecule.
Cells 11 00384 g002
Figure 3. A schematic depiction of ITP cycle.
Figure 3. A schematic depiction of ITP cycle.
Cells 11 00384 g003
Figure 4. Chemical structure of contributing molecules for the ITP cycle, namely xanthine, hypoxanthine and inosine 5-triphosphate.
Figure 4. Chemical structure of contributing molecules for the ITP cycle, namely xanthine, hypoxanthine and inosine 5-triphosphate.
Cells 11 00384 g004
Figure 5. Role of ITPase in genomic stability and cancer.
Figure 5. Role of ITPase in genomic stability and cancer.
Cells 11 00384 g005
Table 1. Clinically relevant ITPA mutants/variants and their biological impacts.
Table 1. Clinically relevant ITPA mutants/variants and their biological impacts.
SNP IDVariationClinical SignificanceBiological SignificanceLocation
rs7270101SNPADRPoor splicing efficiencyc.124 + 21A > C
(g.IVS2 + 21A > C)
rs1127354SNPADRReduced expression, stability, catalysisc.94C > A (p.Pro32Thr)
NASNPEncephalopathyAltered substrate specificity, poor solubilityc.532C > T (p.Arg178Cys)
rs13830SNPTuberculosis3′UTR variation, altered mRNA metabolism/translationg.19176G > A
NANonsenseEncephalopathyNonsense RNA-mediated decayc.452G > A (p.Trp151Stop)
rs863225424DuplicationEncephalopathyFrameshift, non-functional proteinc.359_366dupTCAGCACC (p.Gly123Serfs)
NADeletionEncephalopathy1874 bp deletion, frameshift, non-functional proteinc.264-607_295 + 1267del1906
The overall impacts of ITPA mutations or variants are associated with the other diseases and their treatments, such as infantile encephalopathy, cancer chemotherapy, tuberculosis treatment, Hepatitis C treatment, Antiviral treatment etc.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zamzami, M.A. Inosine Triphosphate Pyrophosphatase (ITPase): Functions, Mutations, Polymorphisms and Its Impact on Cancer Therapies. Cells 2022, 11, 384. https://0-doi-org.brum.beds.ac.uk/10.3390/cells11030384

AMA Style

Zamzami MA. Inosine Triphosphate Pyrophosphatase (ITPase): Functions, Mutations, Polymorphisms and Its Impact on Cancer Therapies. Cells. 2022; 11(3):384. https://0-doi-org.brum.beds.ac.uk/10.3390/cells11030384

Chicago/Turabian Style

Zamzami, Mazin A. 2022. "Inosine Triphosphate Pyrophosphatase (ITPase): Functions, Mutations, Polymorphisms and Its Impact on Cancer Therapies" Cells 11, no. 3: 384. https://0-doi-org.brum.beds.ac.uk/10.3390/cells11030384

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