Next Article in Journal
A Novel Approach to Reducing Lung Metastasis in Osteosarcoma: Increasing Cell Stiffness with Carbenoxolone
Next Article in Special Issue
Gene Expression Profiles of Methyltransferases and Demethylases Associated with Metastasis, Tumor Invasion, CpG73 Methylation, and HPV Status in Head and Neck Squamous Cell Carcinoma
Previous Article in Journal
Cellular and Genetic Background of Osteosarcoma
Previous Article in Special Issue
Deregulation of Autophagy and Apoptosis in Patients with Myelodysplastic Syndromes: Implications for Disease Development and Progression
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Clinicopathological Characteristics and Mutational Landscape of APC, HOXB13, and KRAS among Rwandan Patients with Colorectal Cancer

by
Felix Manirakiza
1,2,3,
Eric Rutaganda
4,
Hidetaka Yamada
3,
Yuji Iwashita
3,
Belson Rugwizangoga
1,2,
Benoit Seminega
4,
Vincent Dusabejambo
4,
Gervais Ntakirutimana
2,
Deogratias Ruhangaza
5,
Annette Uwineza
2,6,
Kazuya Shinmura
3 and
Haruhiko Sugimura
3,7,*
1
Department of Pathology, College of Medicine and Health Sciences, University of Rwanda, Kigali P.O. Box 3286, Rwanda
2
Department of Pathology, University Teaching Hospital of Kigali, Kigali P.O. Box 655, Rwanda
3
Department of Tumor Pathology, Hamamatsu University School of Medicine, 1-20-1 Handayama, Higashi-Ku, Shizuoka 431-3192, Japan
4
Department of Internal Medicine, University Teaching Hospital of Kigali, Kigali P.O. Box 655, Rwanda
5
Department of Pathology, Butaro Hospital, Musanze P.O. Box 59, Rwanda
6
Department of Biochemistry, Molecular Biology and Genetics, College of Medicine and Health Sciences, University of Rwanda, Kigali P.O. Box 3286, Rwanda
7
Sasaki Institute Sasaki Foundation, 2-2 Kanda Surugadai, Chiyoda-Ku, Tokyo 101-0062, Japan
*
Author to whom correspondence should be addressed.
Curr. Issues Mol. Biol. 2023, 45(5), 4359-4374; https://doi.org/10.3390/cimb45050277
Submission received: 27 March 2023 / Revised: 24 April 2023 / Accepted: 11 May 2023 / Published: 16 May 2023
(This article belongs to the Special Issue Advances in Molecular Pathogenesis Regulation in Cancer)

Abstract

:
Cancer research in Rwanda is estimated to be less than 1% of the total African cancer research output with limited research on colorectal cancer (CRC). Rwandan patients with CRC are young, with more females being affected than males, and most patients present with advanced disease. Considering the paucity of oncological genetic studies in this population, we investigated the mutational status of CRC tissues, focusing on the Adenomatous polyposis coli (APC), Kirsten rat sarcoma (KRAS), and Homeobox B13 (HOXB13) genes. Our aim was to determine whether there were any differences between Rwandan patients and other populations. To do so, we performed Sanger sequencing of the DNA extracted from formalin-fixed paraffin-embedded adenocarcinoma samples from 54 patients (mean age: 60 years). Most tumors were located in the rectum (83.3%), and 92.6% of the tumors were low-grade. Most patients (70.4%) reported never smoking, and 61.1% of patients had consumed alcohol. We identified 27 variants of APC, including 3 novel mutations (c.4310_4319delAAACACCTCC, c.4463_4470delinsA, and c.4506_4507delT). All three novel mutations are classified as deleterious by MutationTaster2021. We found four synonymous variants (c.330C>A, c.366C>T, c.513T>C, and c.735G>A) of HOXB13. For KRAS, we found six variants (Asp173, Gly13Asp, Gly12Ala, Gly12Asp, Gly12Val, and Gln61His), the last four of which are pathogenic. In conclusion, here we contribute new genetic variation data and provide clinicopathological information pertinent to CRC in Rwanda.

1. Introduction

Cancer in Africa is underinvestigated and underreported in the scientific literature [1,2,3], especially with respect to cancer genetics and genomics [4,5]. Available data highlight current and future increases in the incidence and mortality related to colorectal cancer (CRC) in African countries [3,6,7], and the Rwandan population is no exception [7].
Globally, CRC accounted for nearly 10% of cancer incidence and 9% of cancer deaths in 2020 and was ranked as the third-most deadly cancer [8,9].
Rwanda is an East African country with a population of 13,246,394 in 2022 [10]. Cancer-related research in this country is estimated to represent less than 1% of the total African cancer research output [1], and little research exists on cancer genetics in Rwanda [5,11,12,13,14,15].
Despite the paucity of CRC research, the few published data show that, in contrast to other parts of the world where CRC is more frequent in males and/or in older patients [8,16,17], in Rwanda, females are more affected than males [18,19] and many Rwandan CRC patients are young [18,19]. Moreover, most patients are diagnosed with advanced disease, a high number of patients are lost to follow-up, and the overall survival of patients with CRC is poor [18,19]. These effects may be due to the low density of healthcare personnel, limited infrastructure and specialized care, lack of awareness of CRC among the Rwandan population, and the unavailability of screening programs [20]. Furthermore, Awedew et al. [7] reported that Rwanda was among the African countries with the highest percentage change in CRC incidence cases between 2010 and 2019. This period coincides with an improvement in cancer diagnostic capacity within Rwanda as well as the emergence of specialized care. This includes the deployment of more surgeons in remote areas, the availability of chemotherapy, and other factors that may explain the increase in CRC incidence [21,22].
The role of heredity or genetic factors in CRC has been estimated to be approximately 35–40% [23,24]. Moreover, differences in genetic background and molecular characteristics associated with cancer between Africans and other races have also been reported [17,25]. Regarding CRC, differences in microsatellite stability status and somatic mutations have been documented [17,25,26,27]. In particular, the adenomatous polyposis coli (APC) and Kirsten rat sarcoma (KRAS) genes—both of which are among CRC driver genes [27,28]—have been reported to be more altered in patients of African ancestry [26].
APC is a multifunctional tumor suppressor gene [29] and is considered to be a key gatekeeper gene involved in CRC development [28,30]. Schell et al. [30] reported that APC may play a prognostic role in CRC and stated that routine clinical assessment of APC mutations together with KRAS, TP53, and BRAF may be used to predict CRC outcome.
KRAS is the most frequently mutated oncogene in tumor tissues [31]. There are approved drugs targeting KRAS, while others are currently in clinical trials [31]. Recently, however, Zafra et al. [31], using a newly developed preclinical model, found that each allele-specific KRAS “mutation has distinct cellular consequences in vivo and carries differential sensitivity to targeted therapeutic agents”. Furthermore, a higher mutation rate in the KRAS gene has been reported in South African and African-American patients with CRC compared to Caucasians and/or other races [26,27,32,33].
Given the rising number of CRC cases in the Rwandan population, the young mean age of diagnosis, and the scarcity of oncogenetic studies in this population, we conducted this study to evaluate the mutational status in tumor tissues. We focused on the APC and KRAS genes, which are the most frequently mutated CRC driver genes. We also investigated the HOXB13 gene, which has been reported to be underexpressed in distal colon cancer cases [34,35], but for which current data are conflicting regarding the association between its mutational status and CRC risk [34,36,37,38]. Our aim was, therefore, to elucidate any differences between the Rwandan population and other populations. As a result, in this paper, we describe the frequency of genetic variation in the KRAS, HOXB13, and APC genes, including novel mutations in APC in CRC tissues from Rwandan patients.

2. Materials and Methods

2.1. Ethical Approval and Consent to Participate

This study was approved by the Institutional Review Board (IRB) of the University of Rwanda College of Medicine and Health Sciences (Approval notice No. 048/CMHS IRB/2020, No. 113/CMHS IRB/2021, and No. 176/CMHS IRB/2022), the Ethics Committee of the University Teaching Hospital of Kigali (Ref.: EC/CHUK/2/064/2020) and the Ethics Committee of the Hamamatsu University School of Medicine (EC HUSM number: 20-011). All participants were informed of the purpose of the study, and they signed an informed consent form prior to participation in this study. The study was conducted in accordance with the Declaration of Helsinki.

2.2. Patients

From December 2020 to September 2022, we prospectively recruited consecutive cases of participants from patients attending colonoscopy services at the Department of Internal Medicine at the University Teaching Hospital of Kigali (CHUK, French acronym), Rwanda. Clinical data and self-reported information regarding the family history of cancer, alcohol, and tobacco use were recorded prospectively. Tissue processing and microscopic examination were performed at CHUK. The microscopic diagnosis was reconfirmed by a second pathologist at Hamamatsu University School of Medicine (HUSM) in Japan, and only patients with histologically confirmed CRC were included in this study. Initially, we graded cases according to the extent of glandular differentiation/formation using a three-tier grading system as follows: grade 1 (well-differentiated CRC, glandular formation >95%), grade 2 (moderately differentiated CRC, glandular formation 50–95%), and grade 3 (poorly differentiated CRC, glandular formation <50%) [39]. All CRC cases were then grouped into a two-tier grading system: low grade (i.e., including grade 1 and grade 2) and high grade (i.e., including grade 3), according to the recommendations of the fifth edition of the World Health Organization (WHO) classification of tumors of the digestive system [40]. During our study period, 148 consecutive patients underwent colonoscopic biopsy for suspected CRC. Of these, 129 (87.1%) signed an informed consent form to participate in our study. Here, 58 of 129 (44.9%) subjects had CRC, which was confirmed by biopsy, but 4 cases were excluded: one due to poor quality DNA, one due to a low concentration of extracted DNA, one case whose DNA could not be amplified by PCR, and one due to insufficient tissue for DNA extraction. The final number of cases analyzed and presented in this study was 54, and all cases included here were naïve to cancer therapy.

2.3. DNA Extraction

Since our samples were collected prospectively, biopsy specimens were immediately fixed in a 10% neutral buffered formalin solution at the time of biopsy. After automated tissue processing, samples were embedded in paraffin. DNA extraction was performed before the tissues were six months old. For each case, the DNA was extracted from formalin-fixed paraffin-embedded (FFPE) tissue blocks containing at least 50% tumor cells. Extraction was performed using QIAamp DNA FFPE Advanced UNG Kits (Qiagen GmbH, Hilden, Germany), with all procedures following the manufacturer’s recommendations. The concentration and quality of extracted DNA were then measured using a Nanodrop® 1000 spectrophotometer (ThermoFisher Scientific, Wilmington, CO, USA). As measured using the Nanodrop, the cases with a 260/280 ratio between 1.7 and 2.3 were included for further analysis.

2.4. DNA Amplification and Sanger Sequencing

For the APC gene, our sequencing targeted the mutation cluster region (MCR) [41,42,43,44] with flanking bases. The sequenced region was, therefore, GRCh 38: 112838850 to 112840618; this includes the MCR with flanking nucleotides.
We also sequenced both exons (i.e., exons 1 and 2) of the HOXB13 gene [45]. Moreover, to cover a wide range of KRAS mutations in an understudied population—since there are no available genetic studies of Rwandans with CRC—we sequenced exons 2 to 4 of KRAS [46]. We also added a fragment of exon 5 extending from GRCh 38: 25209690 to 25209999 to include the hypervariable region (HVR) of this gene since mutations corresponding to the HVR are not included in some reports [47]. The boundaries of all sequenced exon segments are shown in Supplementary Table S1. Prior to sequencing, extracted DNA was amplified by PCR using HotStarTaq DNA Polymerase (Qiagen) in a 20 µL reaction volume. A list of primer sequences used to amplify the three genes is provided in Supplementary Table S1. The quality of PCR products was then assessed by electrophoresis on a 2% agarose gel. After amplification, PCR products were purified and sequenced as described in a previous study [48]. Briefly, PCR products were purified using ExoSAP-IT (ThermoFisher Scientific). The resulting purified products were directly Sanger sequenced in both directions using the BigDyeTM Terminator v3.1 Cycle Sequencing Kit (Applied BiosystemsTM). This was performed in 20 µL reaction volumes, with all procedures performed according to the manufacturer’s recommendations. Base calling was performed using an Applied Biosystems 3130xl or 3500xL Genetic Analyzer (Applied Biosystems, Foster City, CA, USA).

2.5. TA Cloning and Plasmid Sequence Analysis for Insertion/Deletion Cases

TA cloning was performed to assess cases with insertion/deletion mutations. Note that two separate cloning experiments (i.e., including two different transformations, with two PCR products, obtained in separate reactions) were used to eliminate possible polymerase-induced errors.

2.5.1. Ligation Reaction Using pGEM®-T Easy Vector System

First, PCR products were purified using QIAquick® PCR Purification Kits (Qiagen GmbH, Hilden, Germany). Next, we constructed plasmids by ligating purified PCR products to a pGEM®-T Easy Vector System (Promega Corporation, Madison, WI, USA) with all procedures following the manufacturer’s protocol.

2.5.2. Transformation Using Highly Competent DH5α E. coli Cells

COMPETENT High DH5α competent E. coli cells (Toyobo Corporate, Osaka, Japan) were then transformed with the constructed plasmid. All procedures were performed according to Promega’s protocol, except that we used DH5α E. coli cells (Toyobo). Furthermore, we prepared fresh X-Gal/IPTG Luria-Bertani (LB) agar plates containing ampicillin at a final concentration of 100 µg/µL according to the protocol of ThermoFisher Scientific (i.e., to each 25 mL of LB agar, we added 40 µL of X-Gal solution (20 mg/mL) and 40 µL of IPTG (100 mM); then, the plates were allowed to dry under UV light at room temperature before use).

2.5.3. Screening for Transformants followed by Sanger Sequencing of Positive Subclones

Transformed cells were grown overnight on X-gal/IPTG LB agar plates described above, allowing the selection of blue/white colonies. White colonies (i.e., eight subclones) were then selected [48] and each cultured in 50 µL of liquid LB medium containing ampicillin at a final concentration of 100 µg/µL for 1 h and 30 min at 37 °C. The resulting cultures were centrifuged, and 1 µL of the lower phase was used for PCR in a 20 µL volume using HotStartTaq (Qiagen) polymerase enzyme and pGEM Easy V-2930F (5′-AGG CGA TTA AGT TGG GTA ACG-3′) and pGEM Easy V-134R (5′-CAA GCT ATG CAT CCA ACG C-3′) primers. The resulting PCR products were then purified and sequenced as described above.

2.6. Mutation Detection, Annotation, and In Silico Analysis

ABI sequences were aligned to the respective genomic sequences for each gene (i.e., NC_000005.10 for APC, NC_000017.11 for HOXB13, and NC_000012.12 for KRAS) using both UniproUGENE version 45 [49] and GENETYX® version 14.1.0 (Genetyx Corporation, Tokyo, Japan).
Known variants were then annotated using the dbSNP (build 156) database as per the recommendations of the Human Genome Variation Society (HGVS) [50]. Novel variants were analyzed using MutationTaster2021 [51] in silico tools to predict the consequences of the resulting proteins. Each detected mutation was confirmed by two independent experiments to exclude false mutations. We then consulted major databases (i.e., ALFA, ClinVar, dbSNP, VariSome, 1000 Genome Project, TOPMed, and COSMIC) before concluding that a mutation was novel. Finally, since we used tumor tissue, we could not determine whether mutations were purely somatic or were also present in the germline.

3. Results

3.1. Clinicopathological Characteristics of Patients

In this study, we analyzed data from 54 patients with a mean age of 60 ± 13 years (range 31–89 years). Our dataset included 34 females (63%) and 20 males (37%). We found that tumors were mainly located in the rectum (45 cases, 83.3%) or colon (8 cases, 14.8%) with only one case of anorectal tumor (1.9%). All cases were biopsies and histologically confirmed as adenocarcinoma. In total, 50 cases (92.6%) were low-grade and 4 cases (7.4%) were of high-grade. Most patients (38 cases, 70.4%) reported that they had never smoked, while 33 patients (61.1%) reported that they had consumed alcohol. Detailed clinical data are presented in Table 1.

3.2. Characteristics of APC Genetic Variants

We found 27 different genetic variants (Table 2). These included 3 novel genetic mutations (Figure 1), 23 previously reported mutations, and 1 single-nucleotide polymorphism (SNP), rs41115. This SNP has a minimal allele frequency (MAF) of >50% in major databases. Excluding rs4115, 27 (27/54; 50%) cases were identified to have at least 1 mutation. Here, 1 case had 3 mutations, and 4 cases had 2 different mutations each, yielding a total of 33 mutations. Recurrent mutations included rs74380081 (4/54), rs386833391 (3/54), rs121913224 (2/54), rs139387758 (2/54), and rs587783031 (2/54). Further, 19 of 27 (70.3%) mutations were nonsense or frameshift mutations that led to a premature stop codon and are, therefore, classified as pathogenic or deleterious variants. Seven mutations are benign or likely benign (7/27; 25.9%) and one (1/27, 3.70%) variant was of uncertain significance.

3.3. Characteristics of HOXB13 Genetic Variants

We identified four different genetic variants: rs33993186 (c.330C>A; p.Pro110=), rs8556 (c.366C>T; p.Ser122=), rs9900627 (c.513T>C; p.Ser171=), and rs138675188 (c.735G>A; p.Lys245=), all of which are synonymous variants and were classified as benign or likely benign (Table 3). However, the HOXB13 Ter285Lys variant, which has recently been described in the West African population [52], and the most commonly studied HOXB13 Gly84Glu variant were both absent in our study participants.

3.4. Characteristics of Genetic Variants in the KRAS Gene

Next, analysis of the KRAS gene identified six different genetic variants including four pathogenic variants (i.e., Gly12Ala, Gly12Asp, Gly12Val, and Gln61His), one variant with a conflicting interpretation of pathogenicity (Gly13Asp), and one SNP (NM_004985.5: c.519T>C) that leads to a synonymous amino acid (Asp173=). Detailed information regarding these mutations is provided in Table 4.

4. Discussion

CRC is underinvestigated in the Rwandan population. To our knowledge, only two papers [18,53] and one abstract [19] reporting on CRC in Rwanda have been published in international peer-reviewed journals within the last 10 years. In contrast to global data where CRC is reported to be more common in males than in females [9], in this study, we found that CRC was more common in females than in males (i.e., 63.0% versus 37.0%). Our findings also report a higher rate of female incidence than the 52.5% reported by Uwamariya et al. [18] and the 52.2% reported by Fadelu et al. [19].
Most tumors were located in the rectum (83.3%), which may be due to the fact that rectal tumors are more symptomatic (i.e., they display tenesmus, pain during defecation, bleeding, and blood in the stool, among other symptoms) than colon tumors. Therefore, patients with rectal tumors are more likely to seek medical care relative to patients with colon tumors. In addition, screening programs for CRC are not available in our settings [20]. As a result, most patients with colon tumors present with advanced stages of cancer and signs of obstruction. They, therefore, immediately undergo resection to relieve the obstruction without undergoing colonoscopy. This speculation may be supported by the fact that Uwamariya et al. [18] found that rectal cancer accounted for 40% of all CRC cases in a study conducted at the same institution as ours that used both biopsy and resection specimens.
We also found that 35.2% (19/54) of our research participants reported that they had no information regarding a family history of cancer. This is different from missing information that was not recorded, which accounted for 3.7% (2/54 cases) in this study. In addition, 61.1% of CRC patients reported that they had “no family history of cancer.” Therefore, none of the patients in this study reported having a family history of cancer. We speculate that self-reported data regarding family history of cancer may not necessarily mean that cancers were absent, especially because of limited opportunities to receive a cancer diagnosis, treatment, and cancer registration in the past [20]. Therefore, data on family history of cancer should be interpreted with caution, and prospective studies with good documentation regarding cancer-related information are recommended.

4.1. New Mutations in the APC Gene

The APC gene is a key gatekeeper gene involved in the development of CRC [28,30]. Germline mutations in APC are also key factors in familial adenomatous polyposis syndrome. This syndrome is quasi-nonexistent in African populations, with only a few cases having ever been reported in the literature [54,55,56,57,58,59,60,61,62,63]. Mutations in the APC gene in CRC have been reported in more than 50% of cases [44]. Moreover, >60% of APC mutations are located in the MCR [44,64,65,66]. In this study, we limited our genetic assessment to a region including the MCR with ±300 bases flanking either side. In total, we detected 27 genetic variants or mutations, 24 of which have been previously documented in the dbSNP, ClinVar, and/or COSMIC databases.
Three mutations, i.e., c.4310_4319delAAACACCTCC: p.Lys1437Asnfs*33, c.4463_4470delinsA: p.Leu1488Tyrfs*17, and c.4506_4507delT: p.Ser1503Hisfs*4 have, to our knowledge, not yet been reported in the literature or major genetic variation databases. All these mutations are frameshift mutations that are predicted to cause a premature termination codon and therefore produce a truncated protein. Using in silico bioinformatics tools, each of these genetic variants is predicted to be deleterious and to cause nonsense-mediated mRNA decay.

4.2. Variants in the HOXB13 Gene

We identified four different genetic variants: rs33993186, rs8556, rs9900627, and rs13865188, all of which are synonymous and classified as benign or likely benign. Silva et al. [34] only reported the rs8556 and rs9900627 variants in CRC samples from patients treated at the Portuguese Institute of Oncology-Porto. However, in our study, we found cases containing the rs33993186 and rs138675188 variants. The rs33993186 (g.487282264G>T) variant and the rs138675188 (g.48726910C>T) variant appear to be more than two times more common in Africans than in the global population as currently reported in dbSNP build 156.
The HOXB13 Gly84Glu variant found mostly in the Europeans [67,68], the Gly132Glu and Gly135Glu mutations described in the Asians [69,70], and the HOXB13 Ter285Lys variant found mostly in the West African populations [52] were all absent in our study participants.

4.3. KRAS Missense Mutations

Five of the six KRAS variants identified in this study are missense mutations. These mutations included the pathogenic mutations Gly12Asp, Gly12Ala, Gly12Val, and Gln61His as well as the variant Gly13D, which had conflicting interpretations of pathogenicity. These five missense mutations are the most commonly reported KRAS mutations and were present in 22 of 54 samples (40.7%) analyzed here. Ben Salah et al. [71] reported a KRAS mutation rate of 55.7%, while Marbun et al. [72] reported a rate of 64%. In other studies, either Gly12Asp [73,74] or Gly12Val [71,75] were the most common KRAS mutations. However, in our study, Gly13Asp was the most recurrent mutation, followed by Gly12Val. KRAS Gly12Asp, which has been reported to be associated with the best prognosis [73], was present in 2/54 of our patients, and Gly12Cys, which is associated with a poor prognosis [73], was not identified in any patients in our study.

4.4. Smoking, Alcohol Consumption, Diet, and Other Environmental Risk Factors for CRC

A total of 20% (11/54) of our study population reported ever having smoked and 61.1% (33/54) reported ever having consumed alcohol. Both proportions are higher than the 6.9% (for smokers) and 14.7% (for drinkers) reported by Wismayer et al. [76] in the Ugandan population neighboring Rwanda. In addition, Wismayer et al. [76] found that former smokers and current and former drinkers had a higher risk of developing CRC.
It has been reported that a high intake of red meat, processed meat, sweetened beverages, a diet low in fiber, and a low intake of dairy products [77,78,79,80] are factors associated with an increased risk of developing CRC. Although we did not assess the diet of our study participants, it is worth noting that the Rwandans rarely consume meat, fish, fruit, and dairy products, but their diet may be rich in starchy foods [81,82,83].

4.5. Limitations

Our study has some limitations. We only used tumor tissues. Therefore, we could not confirm whether genetic variations and/or mutations were purely somatic or also present in the germline. By using Sanger sequencing, we were only able to investigate a small number of cases and genes. Furthermore, with respect to the APC gene, our analysis was limited to only a small portion of this gene. In addition, we did not investigate any genes associated with Lynch syndrome, nor did we determine the chromosomal stability status of our study participants. Further studies using next-generation sequencing techniques are recommended to investigate a large number of genes in a larger pool of cases. Our clinical data did not include the cancer stage; therefore, we provided limited pathological characteristics of our patients. Furthermore, information on smoking, alcohol consumption, and family history of cancer was self-reported, and we were not able to make an objective estimate of the amounts consumed. Therefore, we could not estimate an association between these factors and the presence or absence of a particular mutation. Finally, we were not able to perform immunohistochemical staining or functional studies to further investigate the effect of the described mutations on the expression levels of corresponding proteins and/or to confirm the in silico predicted mutational effect.

5. Conclusions

In this study, we contribute new genetic variation data and clinicopathological information relevant to CRC patients in Rwanda. This genetic information does not yet reflect specific environmental (i.e., dietary and other) components. However, additional studies are required to improve the quality of scientific data characterizing CRC cases in the Rwandan population and improve evidence-based management of this disease. The genetic data provided in this paper also represent a valuable resource for the study of CRC in understudied populations.

Supplementary Materials

The following supporting information is submitted as supplementary material and can be downloaded at: https://www.mdpi.com/article/10.3390/cimb45050277/s1. Table S1: List of primer pairs for PCR amplification.

Author Contributions

Conceptualization, F.M., H.Y., Y.I., B.R. and H.S.; Data curation, F.M., E.R. and G.N.; Formal analysis, F.M., H.Y., B.R. and D.R.; Funding acquisition, H.S.; Investigation, F.M., E.R., B.S., V.D. and G.N.; Methodology, F.M., H.Y., Y.I. and A.U.; Project administration, A.U., K.S. and H.S.; Resources, E.R., H.Y., B.S., V.D. and H.S.; Supervision, H.Y., A.U., K.S. and H.S.; Validation, F.M., H.Y. and G.N.; Writing—original draft, F.M. and H.S.; Writing—review & editing, F.M., E.R., B.R., D.R., A.U., K.S. and H.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Japan Society for the Promotion of Science (22H04979); SECOM Science and Technology Foundation (1-11-1) and Smoking Research Foundation (2019T009).

Institutional Review Board Statement

The study was conducted in accordance with the Declaration of Helsinki, and approved by the Institutional Review Board (IRB) of the University of Rwanda College of Medicine and Health Sciences (Approval notice No. 048/CMHS IRB/2020, No. 113/CMHS IRB/2021, and No. 176/CMHS IRB/2022), the Ethics Committee of the University Teaching Hospital of Kigali (Ref.: EC/CHUK/2/064/2020) and the Ethics Committee of Hamamatsu University School of Medicine (EC HUSM number: 20-011).

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

Felix MANIRAKIZA would like to thank Hamamatsu University School of Medicine and the Otsuka Toshimi Scholarship Foundation for providing scholarships during the period of research related to the findings presented in this paper.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

Abbreviations

ALFAAllele frequency aggregator
APCAdenomatous polyposis coli
CHUKUniversity Teaching Hospital of Kigali
CMHSCollege of Medicine and health sciences
COSMICCatalogue of somatic mutation in cancer
DNADeoxyribonucleic acid
HOXB13Homeobox B13
IRBInstitution review board
KRASKirsten rat sarcoma
LMICsLow- and middle-income countries
SNPSingle-nucleotide polymorphism
WHOWorld health organization

References

  1. Kayamba, V.; Mutale, W.; Cassell, H.; Heimburger, D.C.; Shu, X.-O. Systematic Review of Cancer Research Output From Africa, With Zambia as an Example. JCO Glob. Oncol. 2021, 7, 802–810. [Google Scholar] [CrossRef] [PubMed]
  2. Hamdi, Y.; Abdeljaoued-Tej, I.; Zatchi, A.A.; Abdelhak, S.; Boubaker, S.; Brown, J.S.; Benkahla, A. Cancer in Africa: The Untold Story. Front. Oncol. 2021, 11, 650117. [Google Scholar] [CrossRef] [PubMed]
  3. Arhin, N.; Ssentongo, P.; Taylor, M.; Olecki, E.J.; Pameijer, C.; Shen, C.; Oh, J.; Eng, C. Age-standardised incidence rate and epidemiology of colorectal cancer in Africa: A systematic review and meta-analysis. BMJ Open 2022, 12, e052376. [Google Scholar] [CrossRef] [PubMed]
  4. Odedina, F.T.; Rotimi, S. Promoting cancer genomics research in Africa: A roadmap. Nat. Rev. Cancer 2021, 21, 409–410. [Google Scholar] [CrossRef] [PubMed]
  5. Rotimi, S.O.; Rotimi, O.A.; Salhia, B. A Review of Cancer Genetics and Genomics Studies in Africa. Front. Oncol. 2021, 10, 606400. [Google Scholar] [CrossRef] [PubMed]
  6. Arnold, M.; Sierra, M.S.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global patterns and trends in colorectal cancer incidence and mortality. Gut 2017, 66, 683–691. [Google Scholar] [CrossRef]
  7. Awedew, A.F.; Asefa, Z.; Belay, W.B. Burden and trend of colorectal cancer in 54 countries of Africa 2010-2019: A systematic examination for Global Burden of Disease. BMC Gastroenterol. 2022, 22, 204. [Google Scholar] [CrossRef]
  8. Xi, Y.; Xu, P. Global colorectal cancer burden in 2020 and projections to 2040. Transl. Oncol. 2021, 14, 101174. [Google Scholar] [CrossRef]
  9. Morgan, E.; Arnold, M.; Gini, A.; Lorenzoni, V.; Cabasag, C.J.; Laversanne, M.; Vignat, J.; Ferlay, J.; Murphy, N.; Bray, F. Global burden of colorectal cancer in 2020 and 2040: Incidence and mortality estimates from GLOBOCAN. Gut 2023, 72, 338–344. [Google Scholar] [CrossRef]
  10. National Institute of Statistics of Rwanda (NISR). The Fifth Rwanda Population and Housing Census, Main Indicators Report, February 2023; National Institute of Statistics of Rwanda (NISR): Kigali, Rwanda, 2023; p. 3. ISBN 978-99977-43-09-1.
  11. Habyarimana, T.; Bakri, Y.; Mugenzi, P.; Mazarati, J.B.; Attaleb, M.; El Mzibri, M. Association between glutathione peroxidase 1 codon 198 variant and the occurrence of breast cancer in Rwanda. Mol. Genet. Genom. Med. 2018, 6, 268–275. [Google Scholar] [CrossRef]
  12. Habyarimana, T.; Attaleb, M.; Mugenzi, P.; Mazarati, J.B.; Bakri, Y.; El Mzibri, M. Association of p53 Codon 72 Polymorphism with Breast Cancer in a Rwandese Population. Pathobiology 2018, 85, 186–191. [Google Scholar] [CrossRef]
  13. Habyarimana, T.; Attaleb, M.; Mugenzi, P.; Mazarati, J.B.; Bakri, Y.; El Mzibri, M. CHEK2 Germ Line Mutations are Lacking among Familial and Sporadic Breast Cancer Patients in Rwanda. Asian Pac. J. Cancer Prev. 2018, 19, 375–379. [Google Scholar] [CrossRef]
  14. Uyisenga, J.P.; Debit, A.; Poulet, C.; Frères, P.; Poncin, A.; Thiry, J.; Mutesa, L.; Jerusalem, G.; Bours, V.; Josse, C. Differences in plasma microRNA content impair microRNA-based signature for breast cancer diagnosis in cohorts recruited from heterogeneous environmental sites. Sci. Rep. 2021, 11, 11698:1–11698:15. [Google Scholar] [CrossRef]
  15. Uyisenga, J.P.; Segers, K.; Lumaka, A.Z.; Mugenzi, P.; Fasquelle, C.; Boujemila, B.; Josse, C.; Mutesa, L.; Bours, V. Screening of germline mutations in young Rwandan patients with breast cancers. Mol. Genet. Genom. Med. 2020, 8, e1500. [Google Scholar] [CrossRef]
  16. Brand, M.; Gaylard, P.; Ramos, J. Colorectal cancer in South Africa: An assessment of disease presentation, treatment pathways and 5-year survival. S. Afr. Med. J. 2018, 108, 118–122. [Google Scholar] [CrossRef]
  17. Alatise, O.I.; Knapp, G.C.; Sharma, A.; Chatila, W.K.; Arowolo, O.A.; Olasehinde, O.; Famurewa, O.C.; Omisore, A.D.; Komolafe, A.O.; Olaofe, O.O.; et al. Molecular and phenotypic profiling of colorectal cancer patients in West Africa reveals biological insights. Nat. Commun. 2021, 12, 6821. [Google Scholar] [CrossRef]
  18. Uwamariya, D.; Ruhangaza, D.; Rugwizangoga, B. Pathological Characteristics, Prognostic Determinants and the Outcome of Patients Diagnosed with Colorectal Adenocarcinoma at the University Teaching Hospital of Kigali. Can. J. Gastroenterol. Hepatol. 2022, 2022, 6608870. [Google Scholar] [CrossRef]
  19. Fadelu, T.; Sebahungu, F.; Diasti, K.; Nguyen, C.; Yeh, T.; Shyirambere, C.; Nkusi, E.; Nsabimana, N.; Ruhangaza, D.; DeBoer, R.J.; et al. Patient characteristics and outcomes of colorectal cancer (CRC) at Butaro Cancer Center of Excellence (BCCOE): Results from a retrospective cohort. J. Clin. Oncol. 2020, 38 (Suppl. 15), e16081. [Google Scholar] [CrossRef]
  20. Rubagumya, F.; Costas-Chavarri, A.; Manirakiza, A.; Murenzi, G.; Uwinkindi, F.; Ntizimira, C.; Rukundo, I.; Mugenzi, P.; Rugwizangoga, B.; Shyirambere, C.; et al. State of Cancer Control in Rwanda: Past, Present, and Future Opportunities. JCO Glob. Oncol. 2020, 6, 1171–1177. [Google Scholar] [CrossRef]
  21. Cancedda, C.; Cotton, P.; Shema, J.; Rulisa, S.; Riviello, R.; Adams, L.V.; Farmer, P.E.; Kagwiza, J.N.; Kyamanywa, P.; Mukamana, D.; et al. Health Professional Training and Capacity Strengthening Through International Academic Partnerships: The First Five Years of the Human Resources for Health Program in Rwanda. Int. J. Health Policy Manag. 2018, 7, 1024–1039. [Google Scholar] [CrossRef]
  22. Stulac, S.; Binagwaho, A.; Tapela, N.M.; Wagner, C.M.; Muhimpundu, M.A.; Ngabo, F.; Nsanzimana, S.; Kayonde, L.; Bigirimana, J.B.; Lessard, A.J.; et al. Capacity building for oncology programmes in sub-Saharan Africa: The Rwanda experience. Lancet. Oncol. 2015, 16, e405–e413. [Google Scholar] [CrossRef]
  23. Lichtenstein, P.; Holm, N.V.; Verkasalo, P.K.; Iliadou, A.; Kaprio, J.; Koskenvuo, M.; Pukkala, E.; Skytthe, A.; Hemminki, K. Environmental and Heritable Factors in the Causation of Cancer—Analyses of Cohorts of Twins from Sweden, Denmark, and Finland. N. Engl. J. Med. 2000, 343, 78–85. [Google Scholar] [CrossRef] [PubMed]
  24. Graff, R.E.; Möller, S.; Passarelli, M.N.; Witte, J.S.; Skytthe, A.; Christensen, K.; Tan, Q.; Adami, H.-O.; Czene, K.; Harris, J.R.; et al. Familial Risk and Heritability of Colorectal Cancer in the Nordic Twin Study of Cancer. Clin. Gastroenterol. Hepatol. 2017, 15, 1256–1264. [Google Scholar] [CrossRef] [PubMed]
  25. Irabor, D.O.; Oluwasola, O.A.; Ogunbiyi, O.J.; Ogun, O.G.; Okolo, C.A.; Melas, M.; Gruber, S.B.; Shi, C.; Raskin, L. Microsatellite Instability Is Common in Colorectal Cancer in Native Nigerians. Anticancer Res. 2017, 37, 2649–2654. [Google Scholar] [CrossRef] [PubMed]
  26. Myer, P.A.; Lee, J.K.; Madison, R.W.; Pradhan, K.; Newberg, J.Y.; Isasi, C.R.; Klempner, S.J.; Frampton, G.M.; Ross, J.S.; Venstrom, J.M.; et al. The Genomics of Colorectal Cancer in Populations with African and European Ancestry. Cancer Discov. 2022, 12, 1282–1293. [Google Scholar] [CrossRef]
  27. Guda, K.; Veigl, M.L.; Varadan, V.; Nosrati, A.; Ravi, L.; Lutterbaugh, J.; Beard, L.; Willson, J.K.V.; Sedwick, W.D.; Wang, Z.J.; et al. Novel recurrently mutated genes in African American colon cancers. Proc. Natl. Acad. Sci. USA 2015, 112, 1149–1154. [Google Scholar] [CrossRef]
  28. Augustus, G.J.; Ellis, N.A. Colorectal Cancer Disparity in African Americans: Risk Factors and Carcinogenic Mechanisms. Am. J. Pathol. 2018, 188, 291–303. [Google Scholar] [CrossRef]
  29. Aoki, K.; Taketo, M.M. Adenomatous polyposis coli (APC): A multi-functional tumor suppressor gene. J. Cell Sci. 2007, 120, 3327–3335. [Google Scholar] [CrossRef]
  30. Schell, M.J.; Yang, M.; Teer, J.K.; Lo, F.Y.; Madan, A.; Coppola, D.; Monteiro, A.N.A.; Nebozhyn, M.V.; Yue, B.; Loboda, A.; et al. A multigene mutation classification of 468 colorectal cancers reveals a prognostic role for APC. Nat. Commun. 2016, 7, 11743. [Google Scholar] [CrossRef]
  31. Zafra, M.P.; Parsons, M.J.; Kim, J.; Alonso-Curbelo, D.; Goswami, S.; Schatoff, E.M.; Han, T.; Katti, A.; Fernandez, M.T.C.; Wilkinson, J.E.; et al. An In Vivo Kras Allelic Series Reveals Distinct Phenotypes of Common Oncogenic Variants. Cancer Discov. 2020, 10, 1654–1671. [Google Scholar] [CrossRef]
  32. Cronjé, L.; Becker, P.J.; Paterson, A.C.; Ramsay, M. Hereditary non-polyposis colorectal cancer is predicted to contribute towards colorectal cancer in young South African blacks. S. Afr. J. Sci. 2009, 105, 68–72. [Google Scholar] [CrossRef]
  33. Wang, D.; Agrawal, R.; Zou, S.; Haseeb, M.A.; Gupta, R. Anatomic location of colorectal cancer presents a new paradigm for its prognosis in African American patients. PLoS ONE 2022, 17, e0271629. [Google Scholar] [CrossRef]
  34. Silva, M.G. The Role of HOXB13 Gene in Colorectal Cancer. Master’s Thesis, Universidade do Porto, Porto, Portugal, 2013. [Google Scholar]
  35. Xie, B.; Bai, B.; Xu, Y.; Liu, Y.; Lv, Y.; Gao, X.; Wu, F.; Fang, Z.; Lou, Y.; Pan, H.; et al. Tumor-suppressive function and mechanism of HOXB13 in right-sided colon cancer. Signal Transduct. Target. Ther. 2019, 4, 51:1–51:14. [Google Scholar] [CrossRef]
  36. Akbari, M.R.; Anderson, L.N.; Buchanan, D.D.; Clendenning, M.; Jenkins, M.A.; Win, A.K.; Hopper, J.L.; Giles, G.G.; Nam, R.; Narod, S.; et al. Germline HOXB13 p.Gly84Glu mutation and risk of colorectal cancer. Cancer Epidemiol. 2013, 37, 424–427. [Google Scholar] [CrossRef]
  37. Cai, Q.; Wang, X.; Li, X.; Gong, R.; Guo, X.; Tang, Y.; Yang, K.; Niu, Y.; Zhao, Y. Germline HOXB13 p.Gly84Glu mutation and cancer susceptibility: A pooled analysis of 25 epidemiological studies with 145,257 participates. Oncotarget 2015, 6, 42312–42321. [Google Scholar] [CrossRef]
  38. Laitinen, V.H.; Wahlfors, T.; Saaristo, L.; Rantapero, T.; Pelttari, L.M.; Kilpivaara, O.; Laasanen, S.-L.; Kallioniemi, A.; Nevanlinna, H.; Aaltonen, L.; et al. HOXB13 G84E mutation in Finland: Population-based analysis of prostate, breast, and colorectal cancer risk. Cancer Epidemiol. Biomark. Prev. 2013, 22, 452–460. [Google Scholar] [CrossRef]
  39. Fletcher, C.D.M. Diagnostic Histopathology of Tumors, 5th ed.; Elsevier: Philadelphia, Pennsylvania, 2019; Volume 1, p. 493. [Google Scholar]
  40. WHO Classification of Tumors Editorial Board. Digestive system tumors. In WHO Classification of Tumours Series, 5th ed.; International Agency for Research on Cancer: Lyon, France, 2019; Volume 1, p. 182. [Google Scholar]
  41. Kohler, E.M.; Derungs, A.; Daum, G.; Behrens, J.; Schneikert, J. Functional definition of the mutation cluster region of adenomatous polyposis coli in colorectal tumours. Hum. Mol. Genet. 2008, 17, 1978–1987. [Google Scholar] [CrossRef]
  42. Sameer, A.S.; Shah, Z.A.; Abdullah, S.; Chowdri, N.A.; Siddiqi, M.A. Analysis of molecular aberrations of Wnt pathway gladiators in colorectal cancer in the Kashmiri population. Hum. Genom. 2011, 5, 441. [Google Scholar] [CrossRef]
  43. Cheadle, J.P.; Krawczak, M.; Thomas, M.W.; Hodges, A.K.; Al-Tassan, N.; Fleming, N.; Sampson, J.R. Different combinations of biallelic APC mutation confer different growth advantages in colorectal tumours. Cancer Res. 2002, 62, 363–366. [Google Scholar]
  44. Christie, M.; Jorissen, R.N.; Mouradov, D.; Sakthianandeswaren, A.; Li, S.; Day, F.; Tsui, C.; Lipton, L.; Desai, J.; Jones, I.T.; et al. Different APC genotypes in proximal and distal sporadic colorectal cancers suggest distinct WNT/β-catenin signalling thresholds for tumourigenesis. Oncogene 2013, 32, 4675–4682. [Google Scholar] [CrossRef]
  45. Brechka, H.; Bhanvadia, R.R.; VanOpstall, C.; Vander Griend, D.J. HOXB13 mutations and binding partners in prostate development and cancer: Function, clinical significance, and future directions. Genes Dis. 2017, 4, 75–87. [Google Scholar] [CrossRef] [PubMed]
  46. Sánchez-Ibarra, H.E.; Reyes-Cortes, L.M.; López-Tavera, E.; Luna-Aguirre, C.M.; Barrera-Saldaña, H.A. Complete Screening of Exons 2, 3, and 4 of KRAS and NRAS Genes Reveals a Higher Number of Clinically Relevant Mutations than Food and Drug Administration Quantitative Polymerase Chain Reaction-Based Commercial Kits. Rev. Investig. Clin. 2020, 72, 337–343. [Google Scholar] [CrossRef] [PubMed]
  47. Pantsar, T. The current understanding of KRAS protein structure and dynamics. Comput. Struct. Biotechnol. J. 2020, 18, 189–198. [Google Scholar] [CrossRef] [PubMed]
  48. Natsume, H.; Szczepaniak, K.; Yamada, H.; Iwashita, Y.; Gędek, M.; Šuto, J.; Ishino, K.; Kasajima, R.; Matsuda, T.; Manirakiza, F.; et al. Non-CpG sites preference in G:C > A:T transition of TP53 in gastric cancer of Eastern Europe (Poland, Romania and Hungary) compared to East Asian countries (China and Japan). Genes Environ. 2023, 45, 1. [Google Scholar] [CrossRef] [PubMed]
  49. Okonechnikov, K.; Golosova, O.; Fursov, M. UGENE team Unipro UGENE: A unified bioinformatics toolkit. Bioinformatics 2012, 28, 1166–1167. [Google Scholar] [CrossRef] [PubMed]
  50. den Dunnen, J.T.; Dalgleish, R.; Maglott, D.R.; Hart, R.K.; Greenblatt, M.S.; McGowan-Jordan, J.; Roux, A.-F.; Smith, T.; Antonarakis, S.E.; Taschner, P.E.M. HGVS Recommendations for the Description of Sequence Variants: 2016 Update. Hum. Mutat. 2016, 37, 564–569. [Google Scholar] [CrossRef]
  51. Steinhaus, R.; Proft, S.; Schuelke, M.; Cooper, D.N.; Schwarz, J.M.; Seelow, D. MutationTaster2021. Nucleic Acids Res. 2021, 49, W446–W451. [Google Scholar] [CrossRef]
  52. Darst, B.F.; Hughley, R.; Pfennig, A.; Hazra, U.; Fan, C.; Wan, P.; Sheng, X.; Xia, L.; Andrews, C.; Chen, F.; et al. A Rare Germline HOXB13 Variant Contributes to Risk of Prostate Cancer in Men of African Ancestry. Eur. Urol. 2022, 81, 458–462. [Google Scholar] [CrossRef]
  53. Musabende, M.; Dusabejambo, V.; Kabakambira, J.D.; Ndayisenga, L.; Page, C. Assessment of the Outcome of Gastrointestinal Neoplasms Diagnosed at Kigali University Teaching Hospital. World Wide J. Multidiscip. Res. Dev. 2021, 7, 36–42. [Google Scholar]
  54. Kiringa, K.S.; Zalzal, M.; Bahizi, K.E.; Hangi, M.S.; Kihemba, K.; Bartels, S.A. Familial Adenomatous Polyposis (FAP): A case observed in eastern Democratic Republic of the Congo. Afr. J. Gastroenterol. Hepatol. 2022, 5, 32–39. [Google Scholar] [CrossRef]
  55. Bojuwoye, M.O.; Olokoba, A.B.; Ogunlaja, O.A.; Agodirin, S.O.; Ibrahim, O.K.; Okonkwo, K.C.; Aliyu, A.M. Familial adenomatous polyposis syndrome with colorectal cancer in two Nigerians: A report of two cases and review of literature. Pan Afr. Med. J. 2018, 30, 6:1–6:5. [Google Scholar] [CrossRef]
  56. McQuaide, J.R.; Stewart, A.W. Familial polyposis of the colon in the Bantu. S. Afr. Med. J. 1972, 46, 1241–1246. [Google Scholar]
  57. Sebbaale, A.K.; Obote, W.W.; Kituuka, O. Familial adenomatous polyposis: A case report. East Cent. Afr. J. Surg. 2004, 9, 54–55. [Google Scholar]
  58. Osuagwu, C.C.; Okafor, O.C.; Ezeome, E.R.; Uche, C.E.; Ememonu, C.; Kesieme, E. Familial adenomatous polyposis with synchronous invasive colonic carcinomas and metastatic jejunal adenocarcinoma in a Nigerian male. Rare Tumors 2010, 2, 189–191. [Google Scholar] [CrossRef]
  59. Alese, O.B.; Irabor, D.O. Adenomatous polyposis coli in an elderly female nigerian. Ghana Med. J. 2009, 43, 139–141. [Google Scholar] [CrossRef]
  60. Wismayer, R. Familial Adenomatous Polyposis Coli in East Africa: A Case Report and Review of the Literature. J. Adv. Med. Med. Res. 2020, 32, 74–80. [Google Scholar] [CrossRef]
  61. Kakembo, N.; Kisa, P.; Sekabira, J.; Ogdzediz, D. Colorectal Polyposis in a 15 Year Old Boy in Uganda-Case Report. East Cent. Afr. J. Surg. 2016, 21, 105–108. [Google Scholar] [CrossRef]
  62. Udofot, S.U.; Ekpo, M.D.; Khalil, M.I. Familial polyposis coli: An unusual case in West Africa. Cent. Afr. J. Med. 1992, 38, 44–48. [Google Scholar]
  63. Alfred, R.; Mills, M. Adenomatous polyposis in a young jamaican male of african descent. West Indian Med. J. 2014, 63, 186–188. [Google Scholar] [CrossRef]
  64. Joseph, R.; Little, P.; Hayes, D.N.; Lee, M.S. Characterization of the number and site of APC mutations in sporadic colorectal cancer. J. Clin. Oncol. 2017, 35, 630. [Google Scholar] [CrossRef]
  65. Miyoshi, Y.; Nagase, H.; Ando, H.; Horii, A.; Ichii, S.; Nakatsuru, S.; Aoki, T.; Miki, Y.; Mori, T.; Nakamura, Y. Somatic mutations of the APC gene in colorectal tumors: Mutation cluster region in the APC gene. Hum. Mol. Genet. 1992, 1, 229–233. [Google Scholar] [CrossRef] [PubMed]
  66. Lüchtenborg, M.; Weijenberg, M.P.; Roemen, G.M.J.M.; de Bruïne, A.P.; van den Brandt, P.A.; Lentjes, M.H.F.M.; Brink, M.; van Engeland, M.; Goldbohm, R.A.; de Goeij, A.F.P.M. APC mutations in sporadic colorectal carcinomas from The Netherlands Cohort Study. Carcinogenesis 2004, 25, 1219–1226. [Google Scholar] [CrossRef] [PubMed]
  67. Xu, J.; Lange, E.M.; Lu, L.; Zheng, S.L.; Wang, Z.; Thibodeau, S.N.; Cannon-Albright, L.A.; Teerlink, C.C.; Camp, N.J.; Johnson, A.M.; et al. HOXB13 is a susceptibility gene for prostate cancer: Results from the International Consortium for Prostate Cancer Genetics (ICPCG). Hum. Genet. 2013, 132, 5–14. [Google Scholar] [CrossRef] [PubMed]
  68. Chen, Z.; Greenwood, C.; Isaacs, W.B.; Foulkes, W.D.; Sun, J.; Zheng, S.L.; Condreay, L.D.; Xu, J. The G84E mutation of HOXB13 is associated with increased risk for prostate cancer: Results from the REDUCE trial. Carcinogenesis 2013, 34, 1260–1264. [Google Scholar] [CrossRef] [PubMed]
  69. Momozawa, Y.; Iwasaki, Y.; Hirata, M.; Liu, X.; Kamatani, Y.; Takahashi, A.; Sugano, K.; Yoshida, T.; Murakami, Y.; Matsuda, K.; et al. Germline Pathogenic Variants in 7636 Japanese Patients With Prostate Cancer and 12 366 Controls. J. Natl. Cancer Inst. 2020, 112, 369–376. [Google Scholar] [CrossRef] [PubMed]
  70. Lin, X.; Qu, L.; Chen, Z.; Xu, C.; Ye, D.; Shao, Q.; Wang, X.; Qi, J.; Chen, Z.; Zhou, F.; et al. A novel Germline mutation in HOXB13 is associated with prostate cancer risk in Chinese men. Prostate 2013, 73, 169–175. [Google Scholar] [CrossRef]
  71. Ben Salah, K.; Emaetig, F.; Jebriel, A.; Shagan, A.; El Mahjoubi, E.; Obaid, E. KRAS Testing for Colorectal Cancer Patients: Our Laboratory Experience in The Libyan National Cancer Institute. Libyan Int. J. Oncol. 2022, 1, 14–21. [Google Scholar]
  72. Marbun, V.M.G.; Erlina, L.; Lalisang, T.J.M. Genomic landscape of pathogenic mutation of APC, KRAS, TP53, PIK3CA, and MLH1 in Indonesian colorectal cancer. PLoS ONE 2022, 17, e0267090. [Google Scholar] [CrossRef]
  73. Koulouridi, A.; Karagianni, M.; Messaritakis, I.; Sfakianaki, M.; Voutsina, A.; Trypaki, M.; Bachlitzanaki, M.; Koustas, E.; Karamouzis, M.V.; Ntavatzikos, A.; et al. Prognostic Value of KRAS Mutations in Colorectal Cancer Patients. Cancers 2022, 14, 3320:1–3320:10. [Google Scholar] [CrossRef]
  74. Abudabous, A.; Drah, M.; Aldehmani, M.; Parker, I.; Alqawi, O. KRAS mutations in patients with colorectal cancer in Libya. Mol. Clin. Oncol. 2021, 15, 1–6. [Google Scholar] [CrossRef]
  75. Mounjid, C.; El Agouri, H.; Mahdi, Y.; Laraqui, A.; Chtati, E.; Ech-charif, S.; Khmou, M.; Bakri, Y.; Souadka, A.; El Khannoussi, B. Assessment of KRAS and NRAS status in metastatic colorectal cancer--Experience of the National Institute of Oncology in Rabat Morocco. Ann. Cancer Res. Ther. 2022, 30, 80–84. [Google Scholar] [CrossRef]
  76. Wismayer, R.; Kiwanuka, J.; Wabinga, H.; Odida, M. Risk Factors for Colorectal Adenocarcinoma in an Indigenous Population in East Africa. Cancer Manag. Res. 2022, 14, 2657–2669. [Google Scholar] [CrossRef]
  77. Wang, L.; Du, M.; Wang, K.; Khandpur, N.; Rossato, S.L.; Drouin-Chartier, J.-P.; Steele, E.M.; Giovannucci, E.; Song, M.; Zhang, F.F. Association of ultra-processed food consumption with colorectal cancer risk among men and women: Results from three prospective US cohort studies. BMJ 2022, 378, e068921:1–e068921:10. [Google Scholar] [CrossRef]
  78. Thanikachalam, K.; Khan, G. Colorectal Cancer and Nutrition. Nutrients 2019, 11, 164. [Google Scholar] [CrossRef]
  79. Bradbury, K.E.; Murphy, N.; Key, T.J. Diet and colorectal cancer in UK Biobank: A prospective study. Int. J. Epidemiol. 2020, 49, 246–258. [Google Scholar] [CrossRef]
  80. Vieira, A.R.; Abar, L.; Chan, D.S.M.; Vingeliene, S.; Polemiti, E.; Stevens, C.; Greenwood, D.; Norat, T. Foods and beverages and colorectal cancer risk: A systematic review and meta-analysis of cohort studies, an update of the evidence of the WCRF-AICR Continuous Update Project. Ann. Oncol. 2017, 28, 1788–1802. [Google Scholar] [CrossRef]
  81. Mashingaidze, N.; Ekesa, B.; Ndayisaba, C.P.; Njukwe, E.; Groot, J.C.J.; Gwazane, M.; Vanlauwe, B. Participatory Exploration of the Heterogeneity in Household Socioeconomic, Food, and Nutrition Security Status for the Identification of Nutrition-Sensitive Interventions in the Rwandan Highlands. Front. Sustain. Food Syst. 2020, 4, 47. [Google Scholar] [CrossRef]
  82. Yanagisawa, A.; Sudo, N.; Amitani, Y.; Caballero, Y.; Sekiyama, M.; Mukamugema, C.; Matsuoka, T.; Imanishi, H.; Sasaki, T.; Matsuda, H. Development and Validation of a Data-Based Food Frequency Questionnaire for Adults in Eastern Rural Area of Rwanda. Nutr. Metab. Insights 2016, 9, 31–42. [Google Scholar] [CrossRef]
  83. Paridaens, A.-M. Rwanda 2021|Comprehensive Food Security and Vulnerability Analysis; National Institute of Statistics of Rwanda (NISR): Kigali, Rwanda, 2021; pp. 41–47.
Figure 1. Novel mutations in the APC gene. (A) mutation c.4310_4319delAAACACCTCC, (B) mutation c.4463_4470delinsA and (C) mutation c.4506-4507delT. Upper electropherograms correspond to the wild-type APC sequence and lower electropherograms correspond to mutations at the same positions.
Figure 1. Novel mutations in the APC gene. (A) mutation c.4310_4319delAAACACCTCC, (B) mutation c.4463_4470delinsA and (C) mutation c.4506-4507delT. Upper electropherograms correspond to the wild-type APC sequence and lower electropherograms correspond to mutations at the same positions.
Cimb 45 00277 g001
Table 1. Clinical characteristics of all included cases.
Table 1. Clinical characteristics of all included cases.
Case NoAge at
Diagnosis
SexSiteGradeEver Smoker (Self-Reported)Ever Alcohol Intake (Self-Reported)Family History
of Cancer (Self-Reported)
APC Mutation #KRAS Mutation
149FRectumGrade 2NoYesNo
259FRectumGrade 2NoNoNoYes
331MRectumGrade 1NoYesNoYesYes
483FAnorectumGrade 2NANANo informationYes 2Yes
547FRectumGrade 3NoYesNoYes
650MRectumGrade 2NoYesNo
789FRectumGrade 1YesYesNo
861FRectumGrade 1NANANoYes 3Yes
966FColonGrade 3NoNoNo Yes
1032FRectumGrade 3NoYesNoYesYes
1176FRectumGrade 3NoYesNo information
1251FRectumGrade 1NoNoNo information
1381MRectumGrade 2YesYesNoYes
1449FRectumGrade 2NoNoNoYes
1572MRectumGrade 1NoYesNo
1663FRectumGrade 2NoYesNo Yes
1759FRectumGrade 2NoYesNoYesYes
1861FColonGrade 1NoNoNoYes 2Yes
1968MRectumGrade 2NoYesNo informationYesYes
2059FRectumGrade 1NoYesNo informationYes
2153MColonGrade 2NoYesNoYesYes
2261MRectumGrade 1YesYesNoYesYes
2359MRectumGrade 2NoYesNo Yes
2464FRectumGrade 1NoYesNo informationYes 2
2563MColonGrade 1NoYesNo Yes
2649MColonGrade 1NoYesNo informationYesYes
2756MRectumGrade 1YesYesNo information Yes
2873FRectumGrade 2NoYesNo information
2977FRectumGrade 2YesYesNo informationYes
3073MRectumGrade 2YesYesNo Yes
3162MRectumGrade 2NAYesNoYesYes
3279FRectumGrade 2NoYesNo information
3350FRectumGrade 2NoNoNo
3467FRectumGrade 1NoNoNoYes 2Yes
3565FRectumGrade 2NoYesNoYes
3645MRectumGrade 1NoYesNo informationYes
3740MRectumGrade 2NANANA
3851MRectumGrade 2NoNoNoYesYes
3957FRectumGrade 1YesNoNo information
4063FRectumGrade 2YesYesNo information
4178FRectumGrade 2YesYesNo information Yes
4238MRectumGrade 1NoNoNo
4345FRectumGrade 1NoNoNo
4461MRectumGrade 2NoNoNo
4563FColonGrade 2NoYesNoYes
4675FRectumGrade 1YesYesNo information Yes
4754FRectumGrade 2YesYesNo informationYes
4836MRectumGrade 2NoNoNo information Yes
4965MRectumGrade 1NoNoNoYes
5068FColonGrade 2NoNoNo
5152FRectumGrade 2NANANA
5264FRectumGrade 2NoYesNo
5374FRectumGrade 2NoYesNoYes
5459FColonGrade 1NoNANo informationYes
NA: not available, #: Superscript numbers in this column indicate the number of APC mutations found in the corresponding case. Variants at rs41115 are not included.
Table 2. Details of APC variants in Rwandan patients with colorectal cancer.
Table 2. Details of APC variants in Rwandan patients with colorectal cancer.
NodbSNP IDGRCh38
Coordinate (NC-000005.10)
NM_000038.6Protein Change
(NP_000029.2)
Codon ChangeClinical
Significance $
Number of Cases with Mutation (N = 54)MAF
ALFA/Global
MAF ALFA/
African
MAF 1000G/
Global
MAF 1000G/
African
1rs121913331g.112838934C>Tc.3340C>Tp.Arg1114TerCGA>TGAPathogenic1NANANANA
2rs386833391g.112839054-112839056delGAAc.3460_3462delGAAp.Glu1157delNA*Benign30.000070.00090.00480.0174
3rs138933660g.112839244A>Cc.3650A>Cp.Asn1217ThrAAT>ACTLikely benign/US10.0000650.0000140.0010.0038
4rs74380081g.112839326A>Gc.3732A>Gp.Glu1244=CAA>CAGBenign40.002460.02820.01420.0537
5NAg.112839451_112839453dupAc.3859dupAp.Ile1287Asnfs*4NA*Deleterious (in silico)1NANANANA
6rs1064794229g.112839514_112839518delTAAAAc.3920_3924delTAAAAp.Ile1307Argfs*6NA*Pathogenic1NANANANA
7rs121913224g.112839515_112839525delAAAGAc.3927_3931delAAAGAp.Glu1309Aspfs*4NA*Pathogenic20.0001ZeroNANA
8NAg.112839534delAc.3940delAp.Arg1314Glyfs*7NA*Deleterious (in silico)1NANANANA
9rs1057517558g.112839549_112839550delCc.3956delCp.Pro1319Leufs*2NA*Pathogenic1NANANANA
10rs1554085480g.112839661C>Ac.4067C>Ap.Ser1356TerTCA>TAAPathogenic1NANANANA
11rs121913328g.112839693C>Tc.4099C>Tp.Glu1367TerCAG>TAGPathogenic1NANANANA
12NAg.112839702A>Tc.4108A>Tp.Lys1370TerAAA>TAAPathogenic1NANANANA
13rs1580645082g.112839783G>Tc.4189G>Tp.Glu1397TerGAG>TAGPathogenic1NANANANA
14rs587782518g.112839810C>Tc.4216C>Tp.Gln1406TerCAG>TAGPathogenic1NANANANA
15NAg.112839904_112839913delAAACACCTCCc.4310_4319delAAACACCTCCp.Lys1437Asnfs*33NA*Deleterious (in silico)1NANANANA
16rs111866410g.112839954A>Gc.4360A>Gp.Lys1454GluAAA>GAABenign10.0003920.00990.00080.0023
17rs387906234g.112839979_112839988delAGc.4393_4394delAGp.Ser1465Trpfs*3NA*Pathogenic1NANANANA
18Rs387906234g.112839979_112839988dupAGc.4393_4394dupAGp.Ser1465Argfs*9NA*Deleterious (in silico)1NANANANA
19NAg.112840012A>Gc.4418A>Gp.Asn1473SerAAT>AGTUS1NANANANA
20rs139387758g.112840014G>Ac.4420G>Ap.Ala1474ThrGCT>ACTBenign20.0003260.01050.00260.0098
21NAg.112840057_112840064delinsAc.4463_4470delinsAp.Leu1488Tyrfs*17NA*Deleterious (in silico)1NANANANA
22rs41115g.112840073G>Ac.4479G>Ap.Thr1493=ACG>ACABenign280.610.530.6650.517
23NAg.112840100_112840101delTc.4506_4507delTp.Ser1503Hisfs*4NA*Deleterious (in silico)1NANANANA
24rs587783031g.112840255_112840260delAc.4666delAp.Thr1556Leufs*9NA*Pathogenic1NANANANA
25rs587783031g.112840255_112840260dupAc.4666dupp.Thr1556Asnfs*3NA*Pathogenic2NANANANA
26rs35634377g.112840487T>Cc.4893T>Cp.Ser1631=AGT>AGCBenign/Likely benign10.000980.01040.00080.0023
27rs1554086285g.112840496_112840500dupGc.4906dupGp.Asp1636Glyfs*2NA*Pathogenic1NANANANA
N: total number of cases, NA: not available, NA*: not applicable, US: uncertain significance, $: We used clinical significance values reported in ClinVar except where we specified that these values were determined using in silico tools (i.e., using MutationTaster2021). Novel mutations are shown in bold font.
Table 3. Details of HOXB13 variants in Rwandan patients with colorectal cancer.
Table 3. Details of HOXB13 variants in Rwandan patients with colorectal cancer.
NodbSNP_IDGRCh38
Coordinate
NM_006361.6Protein Change
NP_006352.2
Codon ChangeClinical
Significance $
Number of Cases with Mutation (N = 54)MAF
ALFA/Global
MAF ALFA/
African
1rs33993186g.48728264G>Tc.330C>Ap.Pro110=CCC>CCAbenign60.004030.0220
2rs8556g.48728228G>Ac.366C>Tp.Ser122=AGC>AGTbenign260.1310.172
3rs9900627g.48728081A>Gc.513T>Cp.Ser171=TCT>TCCbenign200.1040.117
4rs138675188g.48726910C>Tc.735G>Ap.Lys245=AAG>AAALikely benign10.000550.0024
N: total number of cases, $: Clinical significance values reported in ClinVar.
Table 4. Details of KRAS variants in Rwandan patients with colorectal cancer.
Table 4. Details of KRAS variants in Rwandan patients with colorectal cancer.
NodbSNP_IDGRCh38 CoordinateNM_004985.5Protein Change
(NP_004976.2)
Codon ChangeClinical
Significance $
Number of Cases with Mutation (N = 54)MAF
ALFA/
Global
MAF ALFA/
African
1rs121913529g.25245350C>Tc.35G>Ap.Gln12AspGGT>GATPathogenic 20.00001Zero
2rs121913529g.25245350C>Gc.35G>Cp.Gln12AlaGGT>GCTPathogenic/
likely pathogenic
1NANA
3rs121913529g.25245350C>Ac.35G>Tp.Gln12ValGGT>GTTPathogenic 7NANA
4rs112445441g.25245347C>Tc.38G>Ap.Gly13AspGGC>GACC.I11NANA
5rs17851045g.25227341T>Gc.183A>Cp.Gln61HisCAA>CACPathogenic/
likely pathogenic
1NANA
6rs1137282g.25209843A>Gc.519T>Cp.Asp173=GAT>GACBenign190.2139450.1763
$: Clinical significance values reported in ClinVar, C.I: Conflicting interpretation of pathogenicity, N: total number of cases, NA: not available.
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

Manirakiza, F.; Rutaganda, E.; Yamada, H.; Iwashita, Y.; Rugwizangoga, B.; Seminega, B.; Dusabejambo, V.; Ntakirutimana, G.; Ruhangaza, D.; Uwineza, A.; et al. Clinicopathological Characteristics and Mutational Landscape of APC, HOXB13, and KRAS among Rwandan Patients with Colorectal Cancer. Curr. Issues Mol. Biol. 2023, 45, 4359-4374. https://doi.org/10.3390/cimb45050277

AMA Style

Manirakiza F, Rutaganda E, Yamada H, Iwashita Y, Rugwizangoga B, Seminega B, Dusabejambo V, Ntakirutimana G, Ruhangaza D, Uwineza A, et al. Clinicopathological Characteristics and Mutational Landscape of APC, HOXB13, and KRAS among Rwandan Patients with Colorectal Cancer. Current Issues in Molecular Biology. 2023; 45(5):4359-4374. https://doi.org/10.3390/cimb45050277

Chicago/Turabian Style

Manirakiza, Felix, Eric Rutaganda, Hidetaka Yamada, Yuji Iwashita, Belson Rugwizangoga, Benoit Seminega, Vincent Dusabejambo, Gervais Ntakirutimana, Deogratias Ruhangaza, Annette Uwineza, and et al. 2023. "Clinicopathological Characteristics and Mutational Landscape of APC, HOXB13, and KRAS among Rwandan Patients with Colorectal Cancer" Current Issues in Molecular Biology 45, no. 5: 4359-4374. https://doi.org/10.3390/cimb45050277

Article Metrics

Back to TopTop