Next Article in Journal
Oncogenic KRAS-Induced Protein Signature in the Tumor Secretome Identifies Laminin-C2 and Pentraxin-3 as Useful Biomarkers for the Early Diagnosis of Pancreatic Cancer
Next Article in Special Issue
Molecular Characterization of Cancer Associated Fibroblasts in Prostate Cancer
Previous Article in Journal
DNA Methylation of Imprinted Genes KCNQ1, KCNQ1OT1, and PHLDA2 in Peripheral Blood Is Associated with the Risk of Breast Cancer
Previous Article in Special Issue
Prognostic Value of IGF2 mRNA-Binding Protein 3 (IGF2BP3) Intratumoral Expression in Melanoma Patients at the Time of Diagnosis: Comparative Analysis of RT-qPCR Versus Immunohistochemistry
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Bottom-Up Approach to the Discovery of Clinically Relevant Biomarker Genes: The Case of Colorectal Cancer

1
Department of Biological Sciences, Royal Holloway University of London, Surrey, Egham TW20 0EX, UK
2
Ashford and St Peter’s Hospitals NHS Foundation Trust, Guildford Road, Chertsey, Lyne KT16 0RQ, UK
*
Authors to whom correspondence should be addressed.
Cancers 2022, 14(11), 2654; https://doi.org/10.3390/cancers14112654
Submission received: 6 May 2022 / Revised: 26 May 2022 / Accepted: 26 May 2022 / Published: 27 May 2022

Abstract

:

Simple Summary

This manuscript describes a novel ‘bottom-up’ approach to the discovery of putative biomarkers of cancer. Unlike traditional ‘omics’ and ‘big data’ research approaches, which gather data rather indiscriminately and at a great cost, our approach relies on the knowledge of cell molecular biology. A very small set of known markers are used to ‘train’ the prediction algorithm, which utilizes accessible and affordable means to analyze and extract relevant information, leading directly to the markers at a fraction of the time and cost of the ‘omics’ research. To illustrate the capabilities of the proposed approach, we applied the method to colorectal cancer, one of the most common and thoroughly studied cancers. Our method yielded an extended and validated set of 138 putative molecular biomarkers. We further tested 42 of these genes and showed that 41 mRNAs are differentially expressed as we predicted. Our method offers a widely applicable strategy for cancer marker discovery.

Abstract

Traditional approaches to genome-wide marker discovery often follow a common top-down strategy, where a large scale ‘omics’ investigation is followed by the analysis of functional pathways involved, to narrow down the list of identified putative biomarkers, and to deconvolute gene expression networks, or to obtain an insight into genetic alterations observed in cancer. We set out to investigate whether a reverse approach would allow full or partial reconstruction of the transcriptional programs and biological pathways specific to a given cancer and whether the full or substantially expanded list of putative markers could thus be identified by starting with the partial knowledge of a few disease-specific markers. To this end, we used 10 well-documented differentially expressed markers of colorectal cancer (CRC), analyzed their transcription factor networks and biological pathways, and predicted the existence of 193 new putative markers. Incredibly, the use of a validation marker set of 10 other completely different known CRC markers and the same procedure resulted in a very similar set of 143 predicted markers. Of these, 138 were identical to those found using the training set, confirming our main hypothesis that a much-expanded set of disease markers can be predicted by starting with just a small subset of validated markers. Further to this, we validated the expression of 42 out of 138 top-ranked predicted markers experimentally using qPCR in surgically removed CRC tissues. We showed that 41 out of 42 mRNAs tested have significantly altered levels of mRNA expression in surgically excised CRC tissues. Of the markers tested, 36 have been reported to be associated with aspects of CRC in the past, whilst only limited published evidence exists for another three genes (BCL2, PDGFRB and TSC2), and no published evidence directly linking genes to CRC was found for CCNA1, SHC1 and TGFB3. Whilst we used CRC to test and validate our marker discovery strategy, the reported procedures apply more generally to cancer marker discovery.

1. Introduction

Colorectal cancer (CRC) within the United Kingdom accounts for over 40,000 new cases per annum and is the third most common cancer in the UK amongst both men and women [1]. Worldwide CRC results in over 570,000 deaths annually [2]. The pathogenesis of CRC is based upon both genetic and epigenetic changes that lead to altered expression levels of oncogenes and tumor suppressor genes, ultimately leading to oncogenesis [3]. The three main pathways involved are the chromosomal instability pathway [4], the microsatellite instability pathway [5] and the CpG island methylator phenotype pathway [4]. Other pathways known to correlate with the pathogenesis of CRC relate to adiponectin, interleukin-6 and small non-coding RNAs (reviewed in [6]). CRC survival rates are high in the early stages of the disease but drop sharply in the late stages [7]. CRC remains largely asymptomatic, and, therefore, early detection via screening methods can aid in higher survival rates and reduce the disease burden.

1.1. Detection and Screening in Colorectal Cancer

Physical examination via endoscopy of the colon and rectum remains the ‘gold’ standard for the diagnosis of CRC. It does also provide the option of removing pre-malignant lesions such as adenomas via a less invasive procedure with a lower rate of mortality and morbidity and a lower overall effect on the quality of life. However, colonoscopy is an invasive procedure, which requires a great degree of patient compliance and, generally, population screening programs generate pressure on services. The fecal hemoglobin (f-Hb) immunohistochemistry test (FIT), fecal occult blood (FOB) test and other similar immunochemical diagnostic tests for the presence of blood in the stool provide a useful testing platform for population screening that is more effective and less costly than a multi-target stool DNA (MT-sDNA) epigenetic test for DNA hypermethylation [8]. There are few other molecular markers in routine use for CRC testing except for carcinoembryonic antigen (CEA) [9] or KRAS [10]. Existing DNA hypermethylation tests target syndecan-2 [11], TFPI2 and SDC2 [12,13]. Other DNA methylation markers linked to CRC include SFRP2 [14], VIM [15,16], FBN2 and TCERG1 [17]. Another group of molecular markers associated with CRC include micro-RNAs ([18,19]). Molecular biomarkers play an ever-increasing role in the diagnosis and prognosis of CRC and its resistance to treatment. However, only a relatively limited number of validated marker molecules and molecular tests are available, of which none are suitable for the early non-invasive detection and molecular diagnosis of CRC [6]. Known biomarkers currently used for the diagnosis of CRC and other cancers often have limited sensitivity or specificity. Therefore, there is a current need to discover new molecular biomarkers suitable for the development of reproducible non-invasive or minimally invasive tests for population screening and to enable the efficient detection, screening, diagnosis or progression of CRC.

1.2. The Top-Down Approach Utilized in ‘Omics-Based Marker Discovery

A variety of ‘omics’ studies and other bioinformatics-only investigations are yielding increasingly larger numbers of novel potential markers and molecular targets of CRC (reviewed in [20,21,22]) and of breast cancer [23], lung cancer [24], and of other cancers [25]. These studies usually aim to identify differentially expressed genes (DEG) and/or disease-related proteins using a variety of sequencing, microarrays, bead arrays and qPCR approaches or proteomics-based studies. A common trend, especially in genome-wide DEG studies, is to conduct pathway enrichment or gene co-expression analyses of the typically large collections of genes that show a degree of differential expression [26,27,28] aiming (1) to identify the biologically meaningful enriched pathways involved, (2) to reconstitute transcriptional networks and also (3) to reduce the number of false positives and narrow down the list of potential markers or indirectly validate them by identifying the biological context of the concerted dysregulation. The other reason for such a traditional approach is a potentially large number of false positive hits typically generated with ‘omics’ approaches. Limiting DEGs to a subset of genes belonging to a few enriched biologically relevant pathways validates such subsets of candidate markers and helps to address the unavoidable technical issues of large scale quantitative ‘omics’ approaches by removing false positive candidates.

1.3. Research Hypothesis and Implementation

We hypothesized whether it is possible to reverse the process, i.e., to start with a small number of well-known and validated CRC markers, and use them in a ‘bottom-up’ approach to help select ‘relevant’ transcriptional and functional networks involved, and then use the vast existing knowledge of transcription regulation, of biological pathways and gene co-expression for the given markers, to expand the range of CRC markers. Our approach is summarized in Figure 1. To this end, we used 10 previously reported, well-documented CRC markers as a ‘training set’ to identify the most relevant transcriptional factors, biological pathways and gene co-expression networks and to predict the existence of a large number of other co-regulated functional genes (potential new markers or targets of CRC). We then used a set of 10 other completely different known CRC-associated proteins to validate the devised prediction procedure and the outcome. Then, to experimentally validate our predictions, we tested the top ~one-third of the ranked predicted markers by quantifying their mRNA expression profiles in biopsies retrieved from surgically removed CRC tissues. Here we report a new approach to marker discovery, which we have validated, and which allowed us to predict many clinically relevant CRC markers starting with as little as 10 known CRC-associated proteins. The reported procedure applies more generally to cancer marker discovery.

1.4. Brief Justification and Explanation of the Experimental Approach Used

1.4.1. Transcription Factors

Gene expression is a finely tuned process that is regulated at genetic, epigenetic or transcriptomic levels [29]. Nevertheless, the transcription levels of individual mRNAs provide a major impact on the expression levels of the encoded proteins, and therefore protein function and the physiological state of the cell. The regulation of the transcription of individual genes is mediated by transcriptional complexes formed by RNA polymerase with one or multiple transcription factors (TFs) and other regulatory proteins assembled around promoter sequences and other cis-regulatory elements (enhancers, silencers or operators). Individual TFs may be involved in the regulation of many different genes, and any gene may be affected by many different TFs. For example, human TP53 gene expression is regulated by 51 different transcription factors including p53 (the protein product of the TP53 gene), which regulates 164 different genes, which in turn regulate over 600 other genes [30]. Therefore, activation of p53 will most likely result in the transcriptional activation of all of these targets, albeit to a different degree, because 215 other TFs share some of the gene targets with p53 [30]. It is reasonable to assume that if a gene is transcriptionally upregulated, then at least one of the TFs known to regulate that gene is likely to have contributed to such upregulation. Furthermore, other gene targets of that same TF may also be affected similarly, i.e., upregulated. A simplified transcription hierarchy and the TF-driven marker discovery process are summarized in Figure 2.

1.4.2. Biological Pathways

Many gene products (proteins) belong to the same multi-subunit protein complexes or are involved in the same biological processes or metabolic or functional pathways. It is therefore reasonable to expect that the expression levels of such genes may also undergo coordinated regulation, for example, to maintain metabolic flux within a functional pathway [31]. Such co-regulation of gene expression may indicate the functional relatedness of their proteins, which might belong to the same protein complex or be involved in the same biological process or pathway. Such information was widely explored by the Gene Ontology Consortium [32] and the GenomeNet network (Kyoto Encyclopedia of Genes and Genomes (KEGG) database), [33] to name just a few. The availability of such functional annotations helps explain the molecular principles underlying biological pathways, by linking genes and biological functions [34] and could therefore provide an insight into the function of yet unknown genes mapped to known pathways. It is, therefore, reasonable to assume that if a gene product belonging to a known biological pathway is upregulated, then other functional constituents of that same pathway may also have been affected, which is likely to require a change in the expression levels of multiple genes [35,36]. An example of one such pathway hsa05210 (CRC, Homo sapiens) is illustrated in Figure 3. If such a pathway is affected in CRC, then the expression levels of all constituent genes may be affected. The same will be true for other pathways (not limited to CRC). Therefore, a small number of proven dysregulated markers might be used to find relevant pathways. Then the existing comprehensive knowledge of such pathways may be used to suggest additional proteins and their genes potentially co-regulated within the identified pathways and the biological systems involved, thus driving the marker discovery process (a bottom-up approach).

1.4.3. Gene Co-Expression

The existence of gene co-expression patterns has been acknowledged for decades and used to cluster genes into groups of highly correlated mRNA subsets [37]. Such expression patterns may indicate common expression regulation pathways for the genes involved [38], or the commonality of functional, metabolic or cellular pathways for the encoded proteins [39] or their metabolites [40]. Deciphering such expression patterns has facilitated the growth of studies looking into this topic [41,42]. Different mechanistic algorithms for clustering gene expression patterns have been explored [43,44]. The phenomenon of gene co-expression conservation has been documented for a wide range of species, and useful resources have been created to allow seamless access to gene co-expression repositories [45,46]. Gene co-expression analysis has been used to assist the functional classification of genes that were found to be co-activated [47]. The reverse might also be true—functionally-related proteins may have their genes co-expressed or even display similar gene expression patterns. The co-expression of genes or their co-regulated splicing does not necessarily place them into the same functional pathways, making such selection independent from pathways or TF analyses. Figure 4 provides an example of the available co-expression information for the gene KRAS, which is a known biomarker that is used to determine targeted treatment in CRC. The reliance on gene co-expression is especially important because neither of the two key prediction approaches (TF, pathways) fully account for the degree and the nature of expression regulation (up, down or more complex). However, although gene co-expression is widely observed, the scarcity of the accurate documented gene co-expression knowledge suggests that this step is more suitable for prioritization of the predicted markers rather than for their selection or de-selection.

2. Materials and Methods

2.1. Biomarkers: Training and Validation Sets

These were selected following a systematic review and literature search focused on biomarkers, specifically, genes with a known association with CRC. This search was performed using MESH terms of ‘(Colon OR Rectal OR Colorectal) AND Cancer’ [48]. Twenty known CRC markers (Table S1) were then randomized into two groups of ten biomarkers each by placing them in alphabetical order and then taking alternate biomarkers and placing them into each of the two groups—the ‘training set’ (10 genes) and the ‘validation set’ (the other 10 genes), summarized in Table 1.

2.2. Transcription Factors

Transcription factors known to be involved in the regulation of each marker gene from the training set were located using the search functionality of the TRRUST database [30] and limiting the analysis to Homo sapiens. All identified TFs for all the 10 training set genes were combined in a single list. In the cases where the same TF was found more than once, duplicate entries were ignored. For each such individual TF identified, another search was conducted to identify all their known target genes. All such identified gene targets were combined in a single list. In the cases where the same target was found more than once, multiple entries were ignored.

2.3. Biological Pathways

Biological pathways known to be associated with each marker gene from the training set were located using the search functionality of the KEGG online resource [33], the analysis was limited to Homo sapiens. All such identified pathways for the 10 marker proteins from the training marker set were combined in a single list. In the cases where the same pathway was identified more than once, multiple entries were ignored. For each pathway identified, a list of associated genes was acquired from the NCBI BioSystems platform [49], the lists of known target genes obtainable from NCBI BioSystems were used, following conversion of Gene IDs to KEGG Gene IDs with a ‘Database to Database’ conversion tool from [50]. All such identified genes were combined in a single list of GI identifiers. In the cases where the same gene was found more than once, multiple entries were ignored. At this stage of the analysis, to enable us to exclude top level and common pathways and focus our attention on disease-specific narrow subsets of genes, we excluded all pathways that contained more than 80 genes within them, we also excluded common pathways known to be unrelated to cancer pathogenesis.

2.4. Validation of the Gene Selection Procedure and Further Ranking of the Predicted Genes

All potentially affected genes identified in TF searches and genes identified using pathway searches were combined such that only those genes identified in both predictions were taken forward (Figure 1a). To validate the procedure and the list of predicted putative biomarkers of CRC, the search procedure was repeated using a set of 10 different CRC-associated markers (‘validation set’, Table 1 and Figure 1b). Predicted biomarkers found in both lists were used for further analysis (Figure 1c). The known gene co-expression data (using the GeneMania tool with default settings, [45]) were used to further rank the list of predicted markers to prioritize them prior to the second round of experimental validation.

2.5. Experimental Validation of the Predicted Biomarkers Using Human Tissues

Ethical approval for the study (IRAS ID number 260946) was obtained from the Health Research Authority (HRA) board of London, Brighton & Sussex (7 October 2019) and from the HRA and Health and Care National Board in Wales (29 October 2019). With each patient’s consent, matched pairs of tissue samples (1 g of cancer and 1 g of normal colonic mucosa) were excised from surgically resected CRC specimens from patients undergoing scheduled surgery for CRC at Ashford and St Peter’s Hospitals NHS Foundation Trust, Chertsey, United Kingdom.

2.6. RNA Extraction

To extract RNA, 100 mg of each tissue was homogenized for 3 min in 1 mL of RNAzol (Sigma-Aldrich). Following that, 400 µL of de-ionized water was added to each RNAzol homogenate and the mixtures were then re-suspended and incubated for 30 min at room temperature, followed by centrifugation at 17,000× g for 15 min at 24 °C. One mL of the supernatants was carefully removed, without disturbing any pellets, transferred to fresh microcentrifuge tubes containing 800 µL of 100% isopropanol, mixed by vortexing, and further incubated at room temperature for 30 min. The RNA was then precipitated by centrifugation at 17,000× g for 15 min at 24 °C. The RNA pellets were washed twice with 600 µL of 75% ethanol for 15 min and precipitated by centrifugation at 8000× g for 3 min. Following the complete removal of the washing solution, the RNA pellets were dried at room temperature for 30 min and re-dissolved in 200 µL of RNase free water. The preparations were then vortexed for 30 s and incubated for 30 min at room temperature to fully dissolve the RNAs. RNA concentration and purity are measured using the Nanodrop 8000 (Thermo Fisher, Waltham, MA, USA). Aliquoted preparations were stored at −20 °C until use.

2.7. cDNA Synthesis

cDNA was synthesized using GoScript™ Reverse Transcription System (Promega, Madison, MI, USA) and following the manufacturer’s protocol. Each 20 µL synthesis reaction contained 4 µL of random primers (Promega), 2 µL of GoScript master mix containing reverse transcriptase, buffer and dNTP (Promega), 5 µL of extracted RNA from human tissue and 9 µL of nuclease-free water. The reaction mixtures were incubated for 10 min at 16 °C, followed by 10 min at 25 °C, followed by 60 min at 42 °C, and a final incubation for 15 min at 72 °C in a programmable PCR thermocycler model number 5331 (Eppendorf, Hamburg, Germany). The reaction mixtures were then cooled to 4 °C and diluted by adding 200 µL of de-ionized water. Diluted cDNA preparations were aliquoted and stored at −20 °C.

2.8. RT-PCR Primer Selection

PCR primers were designed using ‘Primer-BLAST’ [51] to amplify common conservative regions or the selected mRNAs, identified using the Clustal Omega database [52]. The RT-PCR product length was set to the 300 bp–600 bp range and the maximum Tm was set to 72 °C. qPCR product length was set to the maximum of 200 bp. All other parameters were left at default settings. The designed primer sequences are listed in Table S2. The selected primer sets were purchased from Sigma-Aldrich and were stored as 20 µmol/mL water solutions in 500 µL aliquots at −20 °C. Primer sets for the housekeeping genes GAPDH and ACTN1 were used as controls in RT-PCR amplifications. GAPDH, RNA18s and RNA28s were used as controls in qPCR amplifications, and designed as described above.

2.9. RT-PCR

PCR thermocycler model 5331 (Eppendorf) was used for all RT-PCR amplifications. Human colorectal adenocarcinoma cell line cDNA from a moderately differentiated Duke’s D carcinoma was obtained from Sigma-Aldrich (Cat. No C80 12022904) and was used for RT-PCR validation of the primers. Each RT-PCR reaction contained 10 µL of 2x REDTaq® ReadyMix™ ready-to-use Taq DNA polymerase mixture (Sigma-Aldrich, St. Louis, MI, USA), 1 µL each of the forward and reverse PCR primers (20 µmol/mL each), 1 µL template cDNA and water to a final reaction volume of 20 µL. The amplification conditions included an initial denaturing step of 2 min at 95 °C, followed by 30 amplification cycles, each consisting of a 30 s denaturing step at 95 °C, 60 s annealing step, 30 s extension step at 72 °C and a final extension of 5 min at 72 °C. In all reactions, the annealing temperature was set to be 5 °C below the primers’ melting temperature (Tm, as specified by the manufacturer, Table S2). All RT-PCR products were analyzed by electrophoresis in 2% agarose gels.

2.10. Real-Time PCR

All primer pairs used for qPCR were tested for amplification efficiency and were found to be at least 85% efficient in all cases. Quantitative PCR (qPCR) was performed with a StepOne Plus instrument (AB Biosystems, Waltham, MA, USA). All amplification reactions were constructed using ‘master mix’ iQ SYBR Green Supermix (BioRad, Hercules, CA, USA) to which reaction-specific components were added. Each reaction total volume was limited to 20 µL and consisted of the following: 2 µL template cDNA, 10 µL of SYBR Green Supermix, 3 µL mixture of forward and reverse primers for the gene being tested and 5 µL de-ionized water. The reactions were assembled on ice in 96-well plates, sealed with a clear micro sealing film (BioRad) and transferred into the StepOne Plus thermocycler. An initial denaturing cycle of 95 °C for 2 min was followed by 45 cycles consisting of denaturing step at 95 °C for 15 s, an annealing step at 64 °C for 30 s and an extension at 72 °C for 30 s. The final extension was at 72 °C for 5 min. A high-resolution melt (HRM) analysis was then performed at the end of each amplification. The HRM start temperature was 72 °C, the final temperature was 95 °C, and a heating rate of 1 °C per step was used with a holding time of 4 s between steps. Following this, all data were analyzed using the StepOne analysis software v2.3 supplied by the manufacturer (AB Biosystems). All predicted biomarkers were tested by qPCR amplification of cDNA from three different patients, and each amplification was performed in triplicate. The cycle threshold (Ct) corresponding to the exponential amplification phase was determined and normalized to the endogenous levels of three reference RNAs (housekeeping gene GAPDH mRNA, 18S and 28S rRNAs). Then ΔCt values were calculated using the Livak method [53] as shown in Equations (1)–(2)
Δ C t   ( C a n c e r ) = C t   ( B i o m a r k e r   c a n c e r ) C t   ( H K G   c a n c e r )
Δ C t   ( M u c o s a ) = C t   ( B i o m a r k e r   m u c o s a ) C t   ( H K G   m u c o s a )
Δ Δ C t = Δ C t   ( C a n c e r ) Δ C t   ( M u c o s a )
The normalized expression values for the individual biomarkers were calculated using Equation (4)
T r a n s c r i p t i o n   u p r e g u l a t i o n = 2 Δ Δ C t
The significance of changes was estimated by calculating p-values using T.TEST function of Excel.

3. Results

3.1. Prediction of Novel CRC Biomarkers and Validation of the Data Analysis Approach

Following the initial entry of the 10 biomarkers (the ‘training set’, Table 1), this returned a total of 182 upstream transcription factors that were known to target the 10 biomarker genes used in the search (Table S3). A total of 1753 target genes were identified as being regulated by the 182 TFs (Table S4). These genes are likely to include the genes involved in the pathogenesis of CRC as well as CRC biomarker genes. To further narrow the list of potential marker genes a similar search strategy was applied in conjunction with pathway information. Interrogation of the KEGG database using the 10 biomarkers from the training set yielded a total of 134 different biological pathways in which these 10 markers are known to be involved. Of these 134 pathways only 121 were also found in the NCBI ‘Biosystems’ platform [49], and with the view of downstream compatibility, only those 121 pathways were considered for further analysis. These included a mixture of top level pathway definitions, such as hsa01100, ‘metabolic pathways’ containing over 1200 genes, as well as other pathways with no relation to CRC or carcinogenesis in general. These were excluded from further analysis. The remaining 13 pathways are listed in Table S5. A total of 538 target genes were identified as being included in these pathways (Table S6). Because neither this pathway-defined nor the TF-driven prediction approach were entirely accurate, and to reduce the high false negative and high false positive rates of marker discovery, we combined the two predictions (Tables S4 and S6) and selected only those genes that were found in both sets (totaling 193 genes, Table S7).
To validate the proposed prediction strategy, which allowed us to expand a list of 10 known markers (‘training set’, Table 1) into a list of 193 putative CRC-related genes of possible significance as CRC biomarkers (Table S7), we repeated the same prediction procedure using the different set of CRC-related genes (the ‘validation set’, Table 1). The same process was followed for the validation set of biomarkers. In summary, 173 upstream TFs were identified using the ‘validation set’ of 10 biomarkers, and a total of 1685 downstream gene targets were identified for these TFs (Table S8). Out of 82 pathways identified using 10 biomarkers from the validation set, eight pathways were selected for further analysis after the removal of top level pathways and other pathways with no relation to carcinogenesis in general or CRC. These eight pathways contained a total of 322 genes (Table S9). The two sets of predictions were combined to yield 143 putative CRC-related genes, found in both sets of TF- and pathway-driven predictions using the ‘validation set’ of initial markers, Table S10. Remarkably, out of the 143 putative CRC biomarkers identified starting with the 10 known markers from the ‘validation set’, 138 were also found in the list of predicted markers generated starting from the 10 completely different initial CRC markers (‘training set’), thus validating the original hypothesis and the research strategy. These 138 genes (Table S11), predicted using either of the two sets of 10 known marker genes and relying on the ‘guilt by association’ approach, represent DEGs and putative novel markers of CRC.
Prior to further experimental analysis of the predicted markers, we also investigated the known gene co-expression networks. Whilst such information is arguably less complete and more complex than the current knowledge of TFs’ and pathways’ molecular biology, it may still be useful by helping the selection of co-regulated genes, not least because co-expression does mean a degree of co-regulation (whilst the reverse may not necessarily be true). Of the 138 predicted biomarkers, 44 were identified as being co-expressed with at least one of the 20 genes from the ‘training’ and ‘validation’ sets based on the information from GeneMania [45]. These genes were prioritized for further experimental validation by qPCR.

3.2. Experimental Validation of the Predicted Putative Biomarkers

To experimentally validate our predictions, we designed PCR primers to quantify the expression levels of the predicted markers, first, in a cancer cell cDNA from moderately differentiated Duke’s D carcinoma colorectal cell line cDNA using RT-PCR, and later by qPCR amplification of cDNA from surgically removed colorectal tumors. Forty-two out of 44 RT-PCR primer sets tested yielded amplified PCR products of expected lengths, as checked on agarose gels (not shown). These candidate genes were further investigated using qPCR (summarized in Table 2).
The first tumor sample was representative of a moderately differentiated adenocarcinoma of the sigmoid colon, T2N0M0, EMVI negative. Twenty-three of the 42 biomarkers showed significantly altered mRNA expression levels compared to a normal colonic mucosa tissue sample from the same patient. These included 17 upregulated and 6 downregulated genes (Figure 5A, Table S12). The second CRC tissue was representative of a moderately differentiated adenocarcinoma of the sigmoid colon, T4N2M0, EMVI positive. Twenty-five of the 42 biomarkers showed significantly increased mRNA levels and two were significantly downregulated (Figure 5B, Table S13). The third CRC tissue was representative of a moderately differentiated adenocarcinoma of the caecum, T3N1M0, EMVI negative. Thirty-two of the 42 biomarkers showed significantly increased mRNA levels and six were significantly downregulated (Figure 5C, Table S14), summarized in Table 3.
Overall, 40 out of 42 mRNAs tested by qPCR showed significantly upregulated mRNA levels in at least one of the three surgical specimens tested. Transcription of a gene coding for ‘Mothers against decapentaplegic homolog 4’ (SMAD4) was significantly downregulated in two out of three patients tested (no significant change in the third tumor specimen). These results prove that (1) our marker mining procedure does indeed predict differentially expressed genes very well and that (2) our marker prediction approach is capable of predicting both upregulated and downregulated marker genes. The qPCR results fully validate 41 of the 42 predicted and tested differentially expressed genes, and endorse our marker discovery strategy, capable of predicting and identifying extended lists of differentially expressed genes, starting with only partial information about a small number of known cancer-associated genes. Furthermore, the new strategy is clearly capable of identifying both upregulated and downregulated genes. Of the 40 upregulated transcripts, seven were upregulated significantly in all three patients tested (BID, CDK2, CDK4, CDK6, IRS2, PDGFRB and SHC1), whilst 13 mRNAs displayed a high diversity of expression levels, having significantly upregulated levels in at least one of the three tumors tested and significantly downregulated the level in at least one other of the three tumors tested. All but one mRNAs tested (cyclin-dependent kinase inhibitor 1, CDKN1A) showed either upregulation or downregulation in at least one of the three CRC specimens tested. No significant expression changes were detected for CDKN1A mRNA in any of the three specimens tested. All in all, the qPCR expression data validated 41 of the 42 tested markers and indicated that a wider than previously thought variety of CRC molecular signatures may exist, which merits further investigation.

4. Discussion

Despite decades-long marker discovery efforts, currently dominated by larger-scale genomics and proteomics initiatives (reviewed in [54,55,56,57]), only a few molecular biomarkers of CRC are currently known and even fewer are clinically approved for medical use (reviewed in [6]) Traditional ‘omics’ technologies are perfectly capable of mechanistic sifting through tens of thousands of expressed genes and are routinely employed to find transcripts or proteins with altered expression patterns correlating with the disease in question. We aimed to explore a more rational approach to disease marker discovery. Our algorithm relies heavily on the existence of fundamental cell and molecular biology data accumulated over the years. A fundamental principle that defines our research approach is the reliance on the ‘guilt by association’ approach to marker predictions. That approach turned out to be correct, as we have validated successfully our prediction procedure and also experimentally confirmed the markers predicted.
The reliance on transcriptional regulation as an independent predictor is justified. TFs often regulate the transcription of multiple genes, and individual genes may be affected by multiple TFs [58,59]. Knowledge of the transcription regulation networks is therefore bound to reveal relevant and possibly co-regulated genes. Complex biological functions also rely on well-coordinated functional protein networks and therefore require the concerted expression of multiple proteins. In simple cases of linear metabolic pathways, adequate regulation might be achieved with a common regulatory mechanism, with one or just a few TFs regulating all relevant genes involved or a master regulator TF [29]. However, the need to maintain transcriptional responsiveness to multiple independent stimuli and the existence of branched pathways unavoidably result in a complex web of interconnected transcription regulation networks not yet directly identifiable from largely incomplete TF databases. The earlier attempts to reduce the complexity of large ‘omics’ datasets led to the development of a rational system describing gene sets and the products of their translation in terms of their molecular function, cellular location or biological processes. The latter relies on the contributions of many well-coordinated regulatory mechanisms in addition to transcriptional regulation and provides orthogonal selection criteria to complement the information gained from the analysis of known TF networks. It is reasonable to expect high false negative and high false positive rates of marker discovery using either of these approaches due to the incompleteness of bioinformatics resources such as TF/Pathways databases, and the complexity of biological functions, which depend on a multitude of genetic and epigenetic elements, and other regulatory factors, and which are responsive to multiple independent functional stimuli. It was, therefore, important to combine two independent prediction tools, such as TF networks and pathways, to reduce prediction errors. To add further assurance that the selected markers are relevant to the disease we utilized gene co-expression information. In summary, our approach provides a more rational marker discovery procedure, it does not require sifting through tens of thousands of expressed transcripts in the hope of finding one or a few genes with altered expression patterns correlating with the disease in question. Our method is expected to yield disease-relevant marker genes straight away, and these are also expected to be differentially expressed. The actual expression levels and their changes in disease do require experimental confirmation, which we also achieved. Our approach is very different from a traditional ‘omics’ workflow, where any detected differences in genes, proteins or their expression patterns would still require thorough experimental validation to prove their relevance to the disease and the meaningfulness of any differences detected. Two findings reported here prove our research strategy right and validate the markers predicted. Firstly, 138 predicted novel markers were identified starting from either of the two completely different sets of 10 ‘seed’ markers (Figure 1 and Table 1). Secondly, 41 out of the 42 selected transcripts tested by qPCR in three colorectal tumor tissues showed significant differences in their mRNA expression levels (Table 3). These results validate fully both the new research strategy and the biomarkers predicted. Interestingly, there were some notable variations between the expression levels of the 41 transcripts in the three patients tested by qPCR (Table 3). This is likely due to molecular differences existing between the three different tumor specimens tested. Whilst further research is currently underway to generate a more comprehensive landscape of these markers’ expression profiles in CRC tumors, the data reported here validated our experimental approach.
Of the 42 markers predicted and tested here, 36 have been reported in the literature in the past to be associated with different aspects of colorectal, gastric or other relevant cancers (summarized in Supplementary Table S15). Of these, 23 markers (BAD, BAX, CDH1, CDK6, CDKN1A, CDKN1B, CTNNB1, EGF, EGFR, FN1, HRAS, IRS2, KRAS, MAPK3, MDM2, MMP2, PDGFC, PIK3CA, PIK3R3, SMAD4, TGFB1, TGFBR2, TLR2) have been covered in the literature very widely, and multiple publications link these, largely through mutations, to many different cancers, including CRC. For example, mutations in KRAS can be seen in up to 50% of cases in CRC [60,61]. The loss of function of PIK3CA can be seen in 30–50% of cases of CRC [61]. Others have been reported to have good potential to be prognostic biomarkers for CRC (CDKN1A, CDKN1B [62,63,64]), stomach cancer (MMP2, [65]) and esophageal cancer (EGFR, PIK3CA, [66,67]). The other 13 out of the 42 predicted and tested markers (BCL2L1, BID, CCND1, CCND2, CCND3, CCNE1, CDK2, CDK4, CREB1, E2F3, FGFR1, STAT1, VEGFC) are implicated with various cancers including some aspects of colorectal or gastric, and although published evidence exists, it is less abundant. Given the above, we believe that obtaining 36 biomarkers (proven to be related to CRC and to other cancers) out of the 42 markers tested, provides clear evidence that our research strategy is perfectly capable of identifying relevant molecular markers.
Only very limited published evidence exists to link CRC with another three out of the 42 of the tested genes (BCL2, PDGFRB and TSC2). Here the connection is largely limited to a very few published reports pointing to a correlation between genetic variations and CRC risk or to a correlation between gene co-expression and malignant phenotypes [68,69]. Interestingly, one of these genes (BCL2) showed significantly decreased levels of mRNA in one tumor and significantly increased mRNA levels in one other of the three tumor samples tested here, whilst the other two genes showed significant and very strong upregulation in three out of three tumors (PDGFRB) or two out of three tumors tested (TSC2). We did not find prior publications to compare our data with and that justifies further investigation into the significance of the gene expression changes detected in a larger cohort of CRC patients.
There are no reports in the literature linking CRC with the remaining three of the marker genes tested (CCNA1, SHC1 and TGFB3). Our results indicate that these three markers showed significant changes at the mRNA level in CRC. In particular, CCNA1 was significantly upregulated in one tumor and significantly downregulated in one other tumor tissue of the three tumors tested. SMAD4 was significantly upregulated in all three tumors tested and TGFB3 was significantly upregulated in one out of the three tumors.
Interestingly, one of the 42 markers tested (CDKN1A, which encodes p21-a cyclin-dependent kinase inhibitor-1A protein) that has been suggested in the past to have a prognostic value for CRC [63,70,71,72,73] did not show significant changes in the level of mRNA in either of the three tumor tissues tested by qPCR. That was unexpected, taking into account that CDKN1A is known to be regulated by 115 different TFs and that it was the likeliest candidate to have an altered expression level, yet this remained the only one of the 42 markers with no significant changes in its expression levels at mRNA level. This was especially surprising as previous reports indicated altered levels of p21/CDKN1A in colorectal cancers [74], although the changes in CDKN1A expressions between CRC and gastric cancers were not consistent [75]. Past reports associated cancer progression with both increased [76] and decreased [74,77] expression of p21 protein. It is also of interest that a potential prognostic role was suggested for CDKN1A (gene polymorphism and SNPs, which are not directly linked to the mRNA expression levels) [64,78]. Protein p21 is involved in the control of cell cycle progression. p21 expression is controlled at transcriptional and post-transcriptional levels, including by p53, RAS, epigenetic factors and a broad range of extracellular signals that affect p21/CDKN1A expression by acting on a multitude of transcription factors and corresponding cis-regulatory elements in the CDKN1A promoter. p21 has been known to act as either a tumor suppressor or an oncogene, to upregulate or downregulate transcription in response to DNA damage, to protect cells from apoptosis or to promote it, depending on its cellular location, immediate molecular environment, its interacting proteins and its posttranslational modifications (reviewed in [79,80]). Such functional duality has been interpreted by some experts as the reason for not using CDKN1A as a marker or a therapeutic target [79]. It might indeed be difficult to devise a therapeutic modality to selectively target p21/CDKN1A oncogenic properties but not its tumor suppressor properties. However, the multifaceted nature of p21/CDKN1A could potentially yield invaluable insights into a multitude of CRC development pathways driven by diverse genetic and epigenetic factors, and therefore its prognostic potential justifies further investigation.

5. Conclusions

This manuscript details a novel ‘bottom-up’ approach to the discovery of clinically relevant biomarkers. Although the principles of the methods originate from the fields of data science and data analytics, the actual tools used are now available through many online portals, to allow access and efficient use by academic and industrial users in the field. Traditional ‘omics’ and ‘big data’ research approaches have been gathering data rather indiscriminately and at a great cost for a few decades. A few success stories have emerged since, yet few useful markers have been discovered over the years. The reported approach relies on the knowledge generated in the past, including by ‘omics’ and ‘ big data’ sciences, and on accessible and affordable means to extract information leading directly to the relevant markers at a fraction of the cost of the ‘omics’ research. To illustrate the capabilities of the proposed approach and to test our hypotheses we applied the method to colorectal cancer, one of the most common and thoroughly studied cancers. Our method yielded an extended and validated set of 138 putative CRC biomarkers. We have further experimentally checked 42 of these and confirmed that 41 are differentially expressed in surgically removed CRC tumor tissues. We also identified a substantial variability in the expression levels of the newly identified markers, which justifies further investigation to discover and characterize new expanded molecular signatures of a diverse range of colorectal tumors. The ability to identify much-expanded sets of cancer markers should ultimately facilitate the discovery of molecular markers suitable for population screening applications and the early detection of cancer by non-invasive or minimally invasive means, e.g., liquid biopsies such as routine blood tests, which are widely accepted as suitable means of diagnostics. The reported biomarker mining approach to identify extended sets of molecular markers is not limited to CRC and offers a widely applicable strategy for biomarker discovery.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cancers14112654/s1, Table S1: A summary of the 20 CRC biomarkers from published sources, Table S2: RT-PCR and qPCR primers used in this investigation, Table S3: The transcription factors identified in the TRRUST database using the 10 genes of the training set, Table S4: The genes identified in the TRRUST database being regulated by the TFs, selected using the 10 genes from the training set, Table S5: A subset of pathways identified using the 10 genes of the training set and selected for further analysis, Table S6: The genes identified as belonging to the 13 pathways, selected using the 10 genes from the training set, Table S7: The 193 biomarkers identified following interrogation of TF and pathways databases, using the 10 genes from the training set, Table S8: The genes identified in the TRRUST database being regulated by the TFs, selected using the 10 genes from the validation set, Table S9: The genes identified as belonging to the 8 pathways, selected using the 10 genes from the validation set, Table S10: The 143 biomarkers identified following interrogation of TF and pathways databases, using the 10 genes from the training set, Table S11: The 138 predicted biomarkers identified following interrogation of TF and pathways databases using the training and, separately, the validation gene sets, Table S12: Expression of 42 selected mRNA tested in the excised CRC (patient 1), Table S13: Expression of 42 selected mRNA tested in the excised CRC (patient 2), Table S14: Expression of 42 selected mRNA tested in the excised CRC (patient 3), Table S15: Summary of the predicted and tested markers.

Author Contributions

Conceptualization, P.N. and M.S.; methodology, F.K., N.S., P.N. and M.S.; investigation, F.K.; writing—original draft preparation, F.K.; writing—review and editing, M.S.; supervision, P.N. and M.S.; project administration, M.S.; funding acquisition, P.N. and M.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Ashford and St Peter’s Hospitals NHS Foundation Trust to F.K.

Institutional Review Board Statement

Ethical approval for the study (IRAS ID number 260946) was obtained from the Health Research Authority (HRA) board of London, Brighton & Sussex (7 October 2019) and from the HRA and Health and Care National Board in Wales (29 October 2019).

Informed Consent Statement

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

Data Availability Statement

All data are showin in Supplementary Materials, uploaded with this paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. UK Cancer Research. Cancer Research UK Bowel Cancer Statistics. Available online: https://www.cancerresearchuk.org/health-professional/cancer-statistics/statistics-by-cancer-type/bowel-cancer (accessed on 31 October 2021).
  2. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef]
  3. Grady, W.M.; Carethers, J.M. Genomic and epigenetic instability in colorectal cancer pathogenesis. Gastroenterology 2008, 135, 1079–1099. [Google Scholar] [CrossRef] [Green Version]
  4. Marmol, I.; Sanchez-de-Diego, C.; Pradilla Dieste, A.; Cerrada, E.; Rodriguez Yoldi, M.J. Colorectal Carcinoma: A General Overview and Future Perspectives in Colorectal Cancer. Int. J. Mol. Sci. 2017, 18, 197. [Google Scholar] [CrossRef] [Green Version]
  5. Boland, C.R.; Goel, A. Microsatellite instability in colorectal cancer. Gastroenterology 2010, 138, 2073–2087.e3. [Google Scholar] [CrossRef]
  6. Kamel, F.; Eltarhoni, K.; Nisar, P.; Soloviev, M. Colorectal Cancer Diagnosis: The Obstacles We Face in Determining a Non-Invasive Test and Current Advances in Biomarker Detection. Cancers 2022, 14, 1889. [Google Scholar] [CrossRef]
  7. UK Cancer Research. Bowel Cancer Survial Statistics. Available online: https://www.cancerresearchuk.org/health-professional/cancer-statistics/statistics-by-cancer-type/bowel-cancer/survival#heading-Three (accessed on 10 February 2022).
  8. Ladabaum, U.; Mannalithara, A. Comparative Effectiveness and Cost Effectiveness of a Multitarget Stool DNA Test to Screen for Colorectal Neoplasia. Gastroenterology 2016, 151, 427–439.e6. [Google Scholar] [CrossRef] [Green Version]
  9. Au, F.C.; Stein, B.; Tang, C.K. Carcinoembryonic antigen levels in colonic lesions. Am. J. Surg. 1986, 151, 61–64. [Google Scholar] [CrossRef]
  10. Tejpar, S.; Celik, I.; Schlichting, M.; Sartorius, U.; Bokemeyer, C.; Van Cutsem, E. Association of KRAS G13D tumor mutations with outcome in patients with metastatic colorectal cancer treated with first-line chemotherapy with or without cetuximab. J. Clin. Oncol. 2012, 30, 3570–3577. [Google Scholar] [CrossRef]
  11. Oh, T.; Kim, N.; Moon, Y.; Kim, M.S.; Hoehn, B.D.; Park, C.H.; Kim, T.S.; Kim, N.K.; Chung, H.C.; An, S. Genome-wide identification and validation of a novel methylation biomarker, SDC2, for blood-based detection of colorectal cancer. J. Mol. Diagn. 2013, 15, 498–507. [Google Scholar] [CrossRef]
  12. Zhang, L.; Dong, L.; Lu, C.; Huang, W.; Yang, C.; Wang, Q.; Wang, Q.; Lei, R.; Sun, R.; Wan, K.; et al. Methylation of SDC2/TFPI2 and Its Diagnostic Value in Colorectal Tumorous Lesions. Front. Mol. Biosci. 2021, 8, 706754. [Google Scholar] [CrossRef]
  13. Zhang, W.; Yang, C.; Wang, S.; Xiang, Z.; Dou, R.; Lin, Z.; Zheng, J.; Xiong, B. SDC2 and TFPI2 Methylation in Stool Samples as an Integrated Biomarker for Early Detection of Colorectal Cancer. Cancer Manag. Res. 2021, 13, 3601–3617. [Google Scholar] [CrossRef] [PubMed]
  14. Müller, H.M.; Oberwalder, M.; Fiegl, H.; Morandell, M.; Goebel, G.; Zitt, M.; Mühlthaler, M.; Ofner, D.; Margreiter, R.; Widschwendter, M. Methylation changes in faecal DNA: A marker for colorectal cancer screening? Lancet 2004, 363, 1283–1285. [Google Scholar] [CrossRef]
  15. Shirahata, A.; Hibi, K. Serum vimentin methylation as a potential marker for colorectal cancer. Anticancer Res. 2014, 34, 4121–4125. [Google Scholar] [PubMed]
  16. Shirahata, A.; Sakuraba, K.; Goto, T.; Saito, M.; Ishibashi, K.; Kigawa, G.; Nemoto, H.; Hibi, K. Detection of vimentin (VIM) methylation in the serum of colorectal cancer patients. Anticancer Res. 2010, 30, 5015–5018. [Google Scholar]
  17. Yi, J.M.; Dhir, M.; Guzzetta, A.A.; Iacobuzio-Donahue, C.A.; Heo, K.; Yang, K.M.; Suzuki, H.; Toyota, M.; Kim, H.M.; Ahuja, N. DNA methylation biomarker candidates for early detection of colon cancer. Tumour Biol. 2012, 33, 363–372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Yang, I.P.; Tsai, H.L.; Miao, Z.F.; Huang, C.W.; Kuo, C.H.; Wu, J.Y.; Wang, W.M.; Juo, S.H.; Wang, J.Y. Development of a deregulating microRNA panel for the detection of early relapse in postoperative colorectal cancer patients. J. Transl. Med. 2016, 14, 108. [Google Scholar] [CrossRef] [Green Version]
  19. Durán-Vinet, B.; Araya-Castro, K.; Calderón, J.; Vergara, L.; Weber, H.; Retamales, J.; Araya-Castro, P.; Leal-Rojas, P. CRISPR/Cas13-Based Platforms for a Potential Next-Generation Diagnosis of Colorectal Cancer through Exosomes Micro-RNA Detection: A Review. Cancers 2021, 13, 4640. [Google Scholar] [CrossRef]
  20. Bradner, J.E.; Hnisz, D.; Young, R.A. Transcriptional Addiction in Cancer. Cell 2017, 168, 629–643. [Google Scholar] [CrossRef] [Green Version]
  21. Rohr, M.; Beardsley, J.; Nakkina, S.P.; Zhu, X.; Aljabban, J.; Hadley, D.; Altomare, D. A merged microarray meta-dataset for transcriptionally profiling colorectal neoplasm formation and progression. Sci. Data 2021, 8, 214. [Google Scholar] [CrossRef]
  22. Alves Martins, B.A.; de Bulhões, G.F.; Cavalcanti, I.N.; Martins, M.M.; de Oliveira, P.G.; Martins, A.M.A. Biomarkers in Colorectal Cancer: The Role of Translational Proteomics Research. Front. Oncol. 2019, 9, 1284. [Google Scholar] [CrossRef]
  23. Mueller, C.; Haymond, A.; Davis, J.B.; Williams, A.; Espina, V. Protein biomarkers for subtyping breast cancer and implications for future research. Expert Rev. Proteom. 2018, 15, 131–152. [Google Scholar] [CrossRef] [PubMed]
  24. Cheung CH, Y.; Juan, H.F. Quantitative proteomics in lung cancer. J. Biomed. Sci. 2017, 24, 37. [Google Scholar] [CrossRef] [PubMed]
  25. Kumar, V.; Ray, S.; Ghantasala, S.; Srivastava, S. An Integrated Quantitative Proteomics Workflow for Cancer Biomarker Discovery and Validation in Plasma. Front. Oncol. 2020, 10, 543997. [Google Scholar] [CrossRef]
  26. Yihang, Y.; Ji, C.; Jue, W.; Ming, X.; Yunpeng, Z.; Peng, S.; Leilei, L. Identification Hub Genes in Colorectal Cancer by Integrating Weighted Gene Co-Expression Network Analysis and Clinical Validation in vivo and vitro. Front. Oncol. 2020, 10, 638. [Google Scholar] [CrossRef]
  27. Sun, Y.; Mironova, V.; Chen, Y.; Lundh EP, F.; Zhang, Q.; Cai, Y.; Vasiliou, V.; Zhang, Y.; Garcia-Milian, R.; Khan, S.A.; et al. Molecular Pathway Analysis Indicates a Distinct Metabolic Phenotype in Women with Right-Sided Colon Cancer. Transl. Oncol. 2020, 13, 42–56. [Google Scholar] [CrossRef]
  28. Ding, X.; Duan, H.; Luo, H. Identification of Core Gene Expression Signature and Key Pathways in Colorectal Cancer. Front. Genet. 2020, 11, 45. [Google Scholar] [CrossRef] [Green Version]
  29. Wilson, S.; Filipp, F.V. A network of epigenomic and transcriptional cooperation encompassing an epigenomic master regulator in cancer. NPJ Syst. Biol. Appl. 2018, 4, 24. [Google Scholar] [CrossRef]
  30. Netbiolab. Transcriptional Regulatory Relationships Unraveled by Sentence-Based Text Mining. Available online: https://www.grnpedia.org/trrust/ (accessed on 1 December 2019).
  31. Planqué, R.; Hulshof, J.; Teusink, B.; Hendriks, J.C.; Bruggeman, F.J. Maintaining maximal metabolic flux by gene expression control. PLoS Comput. Biol. 2018, 14, e1006412. [Google Scholar] [CrossRef] [Green Version]
  32. Alliance of Genome Resources. The Gene Ontology Resource. Available online: http://geneontology.org/ (accessed on 20 March 2022).
  33. KEGG. Kyoto Encyclopedia of Genes and Genomes. Available online: https://www.genome.jp/kegg/ (accessed on 1 December 2019).
  34. Khatri, P.; Sirota, M.; Butte, A.J. Ten years of pathway analysis: Current approaches and outstanding challenges. PLoS Comput. Biol. 2012, 8, e1002375. [Google Scholar] [CrossRef]
  35. Berlanda, N.; Somigliana, E.; Frattaruolo, M.P.; Buggio, L.; Dridi, D.; Vercellini, P. Surgery versus hormonal therapy for deep endometriosis: Is it a choice of the physician? Eur. J. Obstet. Gynecol. Reprod. Biol. 2017, 209, 67–71. [Google Scholar] [CrossRef]
  36. Creixell, P.; Reimand, J.; Haider, S.; Wu, G.; Shibata, T.; Vazquez, M.; Mustonen, V.; Gonzalez-Perez, A.; Pearson, J.; Sander, C.; et al. Pathway and network analysis of cancer genomes. Nat. Methods 2015, 12, 615–621. [Google Scholar] [CrossRef] [PubMed]
  37. Eisen, M.B.; Spellman, P.T.; Brown, P.O.; Botstein, D. Cluster analysis and display of genome-wide expression patterns. Proc. Natl. Acad. Sci. USA 1998, 95, 14863–14868. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Cramer, P. Organization and regulation of gene transcription. Nature 2019, 573, 45–54. [Google Scholar] [CrossRef] [PubMed]
  39. DeRisi, J.L.; Iyer, V.R.; Brown, P.O. Exploring the metabolic and genetic control of gene expression on a genomic scale. Science 1997, 278, 680–686. [Google Scholar] [CrossRef] [Green Version]
  40. Hughes, T.R.; Marton, M.J.; Jones, A.R.; Roberts, C.J.; Stoughton, R.; Armour, C.D.; Bennett, H.A.; Coffey, E.; Dai, H.; He, Y.D.; et al. Functional discovery via a compendium of expression profiles. Cell 2000, 102, 109–126. [Google Scholar] [CrossRef] [Green Version]
  41. Reel, P.S.; Reel, S.; Pearson, E.; Trucco, E.; Jefferson, E. Using machine learning approaches for multi-omics data analysis: A review. Biotechnol. Adv. 2021, 49, 107739. [Google Scholar] [CrossRef]
  42. Joshi, A.; Rienks, M.; Theofilatos, K.; Mayr, M. Systems biology in cardiovascular disease: A multiomics approach. Nat. Rev. Cardiol. 2021, 18, 313–330. [Google Scholar] [CrossRef]
  43. Yu, G.; Wang, L.G.; Han, Y.; He, Q.Y. clusterProfiler: An R package for comparing biological themes among gene clusters. Omics 2012, 16, 284–287. [Google Scholar] [CrossRef]
  44. Li, Y.; Bie, R.; Teran Hidalgo, S.J.; Qin, Y.; Wu, M.; Ma, S. Assisted gene expression-based clustering with AWNCut. Stat. Med. 2018, 37, 4386–4403. [Google Scholar] [CrossRef]
  45. Ontario Genomics Institute Genome Canada. GeneMania. Available online: https://genemania.org/ (accessed on 1 December 2019).
  46. Obayashi, T.; Kagaya, Y.; Aoki, Y.; Tadaka, S.; Kinoshita, K. COXPRESdb v7: A gene coexpression database for 11 animal species supported by 23 coexpression platforms for technical evaluation and evolutionary inference. Nucleic Acids Res. 2019, 47, D55–D62. [Google Scholar] [CrossRef]
  47. Van Dam, S.; Vosa, U.; van der Graaf, A.; Franke, L.; de Magalhaes, J.P. Gene co-expression analysis for functional classification and gene-disease predictions. Brief Bioinform. 2018, 19, 575–592. [Google Scholar] [CrossRef] [PubMed]
  48. PubMed. NCBI Pub Med. Available online: https://pubmed.ncbi.nlm.nih.gov/ (accessed on 6 May 2022).
  49. Biosystems. NCBI Biosystems. Available online: https://www.ncbi.nlm.nih.gov/biosystems/ (accessed on 31 October 2019).
  50. bioDBnet. Biological Database Network. Available online: https://biodbnet-abcc.ncifcrf.gov/db/db2db.php (accessed on 25 May 2022).
  51. NCBI. NCBI Primer Blast Tool. Available online: https://www.ncbi.nlm.nih.gov/tools/primer-blast/ (accessed on 31 October 2019).
  52. EMBL-EBI. Clusta lOmega. Available online: https://www.ebi.ac.uk/Tools/msa/clustalo/ (accessed on 1 May 2019).
  53. Schmittgen, T.D.; Livak, K.J. Analyzing real-time PCR data by the comparative C(T) method. Nat. Protoc. 2008, 3, 1101–1108. [Google Scholar] [CrossRef] [PubMed]
  54. Wang, K.; Huang, C.; Nice, E.C. Proteomics, genomics and transcriptomics: Their emerging roles in the discovery and validation of colorectal cancer biomarkers. Expert Rev. Proteom. 2014, 11, 179–205. [Google Scholar] [CrossRef] [PubMed]
  55. Chauvin, A.; Boisvert, F.M. Clinical Proteomics in Colorectal Cancer, a Promising Tool for Improving Personalised Medicine. Proteomes 2018, 6, 49. [Google Scholar] [CrossRef] [Green Version]
  56. Li, J.; Ma, X.; Chakravarti, D.; Shalapour, S.; DePinho, R.A. Genetic and biological hallmarks of colorectal cancer. Genes Dev. 2021, 35, 787–820. [Google Scholar] [CrossRef] [PubMed]
  57. Lindhorst, P.H.; Hummon, A.B. Proteomics of Colorectal Cancer: Tumors, Organoids, and Cell Cultures—A Minireview. Front. Mol. Biosci. 2020, 7, 604492. [Google Scholar] [CrossRef]
  58. Casamassimi, A.; Ciccodicola, A. Transcriptional Regulation: Molecules, Involved Mechanisms, and Misregulation. Int. J. Mol. Sci. 2019, 20, 1281. [Google Scholar] [CrossRef] [Green Version]
  59. Lambert, S.A.; Jolma, A.; Campitelli, L.F.; Das, P.K.; Yin, Y.; Albu, M.; Chen, X.; Taipale, J.; Hughes, T.R.; Weirauch, M.T. The Human Transcription Factors. Cell 2018, 172, 650–665. [Google Scholar] [CrossRef] [Green Version]
  60. Chang, Y.Y.; Lin, J.K.; Lin, T.C.; Chen, W.S.; Jeng, K.J.; Yang, S.H.; Wang, H.S.; Lan, Y.T.; Lin, C.C.; Liang, W.Y.; et al. Impact of KRAS mutation on outcome of patients with metastatic colorectal cancer. Hepatogastroenterology 2014, 61, 1946–1953. [Google Scholar]
  61. Ogino, S.; Nosho, K.; Kirkner, G.J.; Shima, K.; Irahara, N.; Kure, S.; Chan, A.T.; Engelman, J.A.; Kraft, P.; Cantley, L.C.; et al. PIK3CA mutation is associated with poor prognosis among patients with curatively resected colon cancer. J. Clin. Oncol. 2009, 27, 1477–1484. [Google Scholar] [CrossRef]
  62. Watanabe, T.; Wu, T.T.; Catalano, P.J.; Ueki, T.; Satriano, R.; Haller, D.G.; Benson, A.B., 3rd; Hamilton, S.R. Molecular predictors of survival after adjuvant chemotherapy for colon cancer. N. Engl. J. Med. 2001, 344, 1196–1206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Belt, E.J.; Brosens, R.P.; Delis-van Diemen, P.M.; Bril, H.; Tijssen, M.; van Essen, D.F.; Heymans, M.W.; Beliën, J.A.; Stockmann, H.B.; Meijer, S.; et al. Cell cycle proteins predict recurrence in stage II and III colon cancer. Ann. Surg. Oncol. 2012, 19, S682–S692. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Cacina, C.; Yaylim-Eraltan, I.; Arikan, S.; Saglam, E.K.; Zeybek, U.; Isbir, T. Association between CDKN1A Ser31Arg and C20T gene polymorphisms and colorectal cancer risk and prognosis. In Vivo 2010, 24, 179–183. [Google Scholar] [PubMed]
  65. Shen, W.; Xi, H.; Wei, B.; Chen, L. The prognostic role of matrix metalloproteinase 2 in gastric cancer: A systematic review with meta-analysis. J. Cancer Res. Clin. Oncol. 2014, 140, 1003–1009. [Google Scholar] [CrossRef] [PubMed]
  66. Yokota, T.; Serizawa, M.; Hosokawa, A.; Kusafuka, K.; Mori, K.; Sugiyama, T.; Tsubosa, Y.; Koh, Y. PIK3CA mutation is a favorable prognostic factor in esophageal cancer: Molecular profile by next-generation sequencing using surgically resected formalin-fixed, paraffin-embedded tissue. BMC Cancer 2018, 18, 826. [Google Scholar] [CrossRef] [PubMed]
  67. Jiang, D.; Li, X.; Wang, H.; Shi, Y.; Xu, C.; Lu, S.; Huang, J.; Xu, Y.; Zeng, H.; Su, J.; et al. The prognostic value of EGFR overexpression and amplification in Esophageal squamous cell Carcinoma. BMC Cancer 2015, 15, 377. [Google Scholar] [CrossRef] [Green Version]
  68. Zheng, L.; Zhan, Y.; Lu, J.; Hu, J.; Kong, D. A prognostic predictive model constituted with gene mutations of APC, BRCA2, CDH1, SMO, and TSC2 in colorectal cancer. Ann. Transl. Med. 2021, 9, 680. [Google Scholar] [CrossRef]
  69. Slattery, M.L.; Herrick, J.S.; Lundgreen, A.; Fitzpatrick, F.A.; Curtin, K.; Wolff, R.K. Genetic variation in a metabolic signaling pathway and colon and rectal cancer risk: mTOR, PTEN, STK11, RPKAA1, PRKAG2, TSC1, TSC2, PI3K and Akt1. Carcinogenesis 2010, 31, 1604–1611. [Google Scholar] [CrossRef] [Green Version]
  70. Zlobec, I.; Baker, K.; Terracciano, L.M.; Lugli, A. RHAMM, p21 combined phenotype identifies microsatellite instability-high colorectal cancers with a highly adverse prognosis. Clin. Cancer Res. 2008, 14, 3798–3806. [Google Scholar] [CrossRef] [Green Version]
  71. Mitomi, H.; Ohkura, Y.; Fukui, N.; Kanazawa, H.; Kishimoto, I.; Nakamura, T.; Yokoyama, K.; Sada, M.; Kobayashi, K.; Tanabe, S.; et al. P21WAF1/CIP1 expression in colorectal carcinomas is related to Kras mutations and prognosis. Eur. J. Gastroenterol. Hepatol. 2007, 19, 883–889. [Google Scholar] [CrossRef]
  72. Bukholm, I.K.; Nesland, J.M. Protein expression of p53, p21 (WAF1/CIP1), bcl-2, Bax, cyclin D1 and pRb in human colon carcinomas. Virchows Arch. 2000, 436, 224–228. [Google Scholar] [CrossRef] [PubMed]
  73. Rau, B.; Sturm, I.; Lage, H.; Berger, S.; Schneider, U.; Hauptmann, S.; Wust, P.; Riess, H.; Schlag, P.M.; Dörken, B.; et al. Dynamic expression profile of p21WAF1/CIP1 and Ki-67 predicts survival in rectal carcinoma treated with preoperative radiochemotherapy. J. Clin. Oncol. 2003, 21, 3391–3401. [Google Scholar] [CrossRef]
  74. Ogino, S.; Nosho, K.; Shima, K.; Baba, Y.; Irahara, N.; Kirkner, G.J.; Hazra, A.; De Vivo, I.; Giovannucci, E.L.; Meyerhardt, J.A.; et al. p21 expression in colon cancer and modifying effects of patient age and body mass index on prognosis. Cancer Epidemiol. Biomark. Prev. 2009, 18, 2513–2521. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Bednarz-Misa, I.; Fortuna, P.; Diakowska, D.; Jamrozik, N.; Krzystek-Korpacka, M. Distinct Local and Systemic Molecular Signatures in the Esophageal and Gastric Cancers: Possible Therapy Targets and Biomarkers for Gastric Cancer. Int. J. Mol. Sci. 2020, 21, 4509. [Google Scholar] [CrossRef] [PubMed]
  76. Singh, P.; Blatt, A.; Feld, S.; Zohar, Y.; Saadi, E.; Barki-Harrington, L.; Hammond, E.; Ilan, N.; Vlodavsky, I.; Chowers, Y.; et al. The Heparanase Inhibitor PG545 Attenuates Colon Cancer Initiation and Growth, Associating with Increased p21 Expression. Neoplasia 2017, 19, 175–184. [Google Scholar] [CrossRef]
  77. Li, Z.; Qiu, R.; Qiu, X.; Tian, T. SNHG6 Promotes Tumor Growth via Repression of P21 in Colorectal Cancer. Cell Physiol. Biochem. 2018, 49, 463–478. [Google Scholar] [CrossRef]
  78. Kang, B.W.; Jeon, H.S.; Chae, Y.S.; Lee, S.J.; Park, J.Y.; Choi, J.E.; Park, J.S.; Choi, G.S.; Kim, J.G. Association between GWAS-identified genetic variations and disease prognosis for patients with colorectal cancer. PLoS ONE 2015, 10, e0119649. [Google Scholar] [CrossRef]
  79. Kreis, N.N.; Louwen, F.; Yuan, J. The Multifaceted p21 (Cip1/Waf1/CDKN1A) in Cell Differentiation, Migration and Cancer Therapy. Cancers 2019, 11, 1220. [Google Scholar] [CrossRef] [Green Version]
  80. Al Bitar, S.; Gali-Muhtasib, H. The Role of the Cyclin Dependent Kinase Inhibitor p21(cip1/waf1) in Targeting Cancer: Molecular Mechanisms and Novel Therapeutics. Cancers 2019, 11, 1475. [Google Scholar] [CrossRef] [Green Version]
Figure 1. A summary of the methodology used to expand the range of molecular biomarkers of colorectal cancer (CRC). Panel (A): A training set of 10 known CRC markers is used to interrogate transcription factor (TF) databases and, separately, functional pathways databases. Panel (B): Validating the procedure using a set of 10 completely different markers with the same procedure as in (A). Panel (C): The virtually identical set of novel markers identified in (A) and (B) is further ranked using gene co-expression information to prioritize the most likely putative markers for further validation using quantitative PCR analysis of surgically resected CRC tissues.
Figure 1. A summary of the methodology used to expand the range of molecular biomarkers of colorectal cancer (CRC). Panel (A): A training set of 10 known CRC markers is used to interrogate transcription factor (TF) databases and, separately, functional pathways databases. Panel (B): Validating the procedure using a set of 10 completely different markers with the same procedure as in (A). Panel (C): The virtually identical set of novel markers identified in (A) and (B) is further ranked using gene co-expression information to prioritize the most likely putative markers for further validation using quantitative PCR analysis of surgically resected CRC tissues.
Cancers 14 02654 g001
Figure 2. A hypothetical map illustrating TF network hierarchy. A master transcription factor is highlighted with green background. Other transcription factors are highlighted using blue-filled shapes. A red gene depicts a ‘seed’ marker gene known to be involved in a disease (CRC in our case). The red dashed lines with arrows indicate the approach to discovering potentially co-regulated genes (blue) that share the same upstream transcription factor(s). Identification of a master transcription factor (purple dashed line) may lead to discovery of other relevant TFs and of other genes (purple).
Figure 2. A hypothetical map illustrating TF network hierarchy. A master transcription factor is highlighted with green background. Other transcription factors are highlighted using blue-filled shapes. A red gene depicts a ‘seed’ marker gene known to be involved in a disease (CRC in our case). The red dashed lines with arrows indicate the approach to discovering potentially co-regulated genes (blue) that share the same upstream transcription factor(s). Identification of a master transcription factor (purple dashed line) may lead to discovery of other relevant TFs and of other genes (purple).
Cancers 14 02654 g002
Figure 3. A pathway involved in proteoglycans in cancer (colorectal cancer, Homo sapiens). Reprinted with permission from [33].
Figure 3. A pathway involved in proteoglycans in cancer (colorectal cancer, Homo sapiens). Reprinted with permission from [33].
Cancers 14 02654 g003
Figure 4. Gene co-expression network for KRAS as depicted using data from the Gene Expression Omnibus on the GeneMania platform. The differing thickness in lines relates to the strength of co-expression (thicker lines show stronger co-expression). Yellow lines denote physical interactions, blue lines denote co-expression, orange lines predict co-expression, blue lines co-localization, burgundy genetic interactions and black denotes similar pathways.
Figure 4. Gene co-expression network for KRAS as depicted using data from the Gene Expression Omnibus on the GeneMania platform. The differing thickness in lines relates to the strength of co-expression (thicker lines show stronger co-expression). Yellow lines denote physical interactions, blue lines denote co-expression, orange lines predict co-expression, blue lines co-localization, burgundy genetic interactions and black denotes similar pathways.
Cancers 14 02654 g004
Figure 5. Experimental validation of differential expression of the predicted CRC marker genes in three patients using qPCR. The expression values on a Log(2) scale are shown. Panel (A): moderately differentiated adenocarcinoma of the sigmoid colon, T2N0M0, EMVI negative. Panel (B): A moderately differentiated adenocarcinoma of the sigmoid colon, T4N2M0, EMVI positive. Panel (C): A moderately differentiated adenocarcinoma of the caecum, T3N1M0, EMVI negative. All amplifications were performed in triplicate. Confidence intervals (p = 0.05) are shown as black error bars. Expression values were normalized to the endogenous levels of three reference RNAs (GAPDH mRNA, 18S and 28S rRNAs). Significantly upregulated genes are labelled in red and significantly downregulated genes are shown in blue.
Figure 5. Experimental validation of differential expression of the predicted CRC marker genes in three patients using qPCR. The expression values on a Log(2) scale are shown. Panel (A): moderately differentiated adenocarcinoma of the sigmoid colon, T2N0M0, EMVI negative. Panel (B): A moderately differentiated adenocarcinoma of the sigmoid colon, T4N2M0, EMVI positive. Panel (C): A moderately differentiated adenocarcinoma of the caecum, T3N1M0, EMVI negative. All amplifications were performed in triplicate. Confidence intervals (p = 0.05) are shown as black error bars. Expression values were normalized to the endogenous levels of three reference RNAs (GAPDH mRNA, 18S and 28S rRNAs). Significantly upregulated genes are labelled in red and significantly downregulated genes are shown in blue.
Cancers 14 02654 g005
Table 1. Known CRC biomarkers used in this study.
Table 1. Known CRC biomarkers used in this study.
Training SetValidation Set
BAG1BAX
BCL-2CDH1
CDKN1ACDKN1B
CXCR4EGFR
ERBB2ESR1
KRASMK167
PIK3CAPLAU
PTENTERT
TFGBRIITP53
TYMSVEGF
Table 2. Top 44 of the predicted genes/proteins of significance to CRC.
Table 2. Top 44 of the predicted genes/proteins of significance to CRC.
Predicted GenesMain Functions or Relevant
Molecular Phenomena
CDK2, CDK4, CDK6, CDKN1A, CDKN1B, CCNA1, CCND1, CCND2, CCND3, CCNE1Regulation of cell cycle
EGF, EGFR, FGFR1, HRAS, KDR *, KRAS, PIK3CA, PIK3R3, TGFB1, TGFB3, TGFBR2Cell growth, proliferation, differentiation or embryogenesis, wound healing
BAD, BAX, BCL2, BCL2L1, BIDRegulation of apoptosis and cell death
CREB1, E2F3, SMAD4, STAT1Transcription factors
CDH1, CTNNB1, FN1Cell adhesion, motility and/or shape
PDGFC, PDGFRB, VEGFCGrowth factors and their receptors
MAPK3, SHC1, IRS2Cellular signaling, signal transduction
PTEN *, TSC2Tumor suppressor genes
MMP2Extracellular metalloproteinase
TLR2Immune system regulation
MDM2Ubiquitin-protein ligase
* KDR and PTEN genes were excluded from further qPCR analysis.
Table 3. Expression of the 42 selected mRNA tested in the excised CRC tissues.
Table 3. Expression of the 42 selected mRNA tested in the excised CRC tissues.
Gene 1Patient 1 Patient 2 Patient 3
Expression 2p-ValueExpression 2p-ValueExpression 2p-Value
CCNA10.890 0.43662.343↑↑0.00460.380↓↓0.0088
CCND11.423 0.18433.433↑↑0.00315.627↑↑↑0.0001
CCND21.093 0.77261.060 0.98895.253↑↑↑0.0003
CCND30.320↓↓0.02601.043 0.75059.560↑↑↑↑0.0001
CCNE11.050 0.92743.413↑↑0.00360.800 0.0602
CDK23.480↑↑0.01164.187↑↑↑0.00022.527↑↑0.0014
CDK42.360↑↑0.00173.170↑↑0.00152.380↑↑0.0056
CDK62.330↑↑0.00483.870↑↑0.00196.717↑↑↑0.0001
CDKN1A1.060 0.65501.250 0.24771.333 0.1679
CDKN1B1.7070.01631.177 0.32775.533↑↑↑0.0001
EGF1.160 0.65284.600↑↑↑0.00139.340↑↑↑↑0.0002
EGFR0.390↓↓0.02161.5630.00473.617↑↑0.0001
FGFR12.690↑↑0.00020.713 0.07673.490↑↑0.0017
HRAS0.280↓↓0.01441.157 0.38941.5600.0151
KRAS0.373↓↓0.00971.230 0.39948.313↑↑↑↑0.0002
PIK3CA0.297 0.05761.5570.01789.830↑↑↑↑0.0006
PIK3R35.847↑↑↑0.00081.130 0.194310.26↑↑↑↑0.0005
TGFB11.617 0.07281.9170.00000.143↓↓↓0.0060
TGFB30.807 0.23601.203 0.80632.830↑↑0.0165
TGFBR20.913 0.27900.5670.01145.737↑↑↑0.0008
BAD1.6700.005210.07↑↑↑↑0.00000.210↓↓↓0.0135
BAX0.870 0.49854.483↑↑↑0.00980.387↓↓0.0049
BCL20.543 0.12710.210↓↓↓0.01841.2530.0331
BCL2L12.430↑↑0.01018.863↑↑↑↑0.00001.097 0.3578
BID7.500↑↑↑0.00026.157↑↑↑0.00017.463↑↑↑0.0002
CREB11.563 0.10422.363↑↑0.000813.94↑↑↑↑0.0000
E2F32.600↑↑0.00711.100 0.77034.443↑↑↑0.0012
SMAD40.5070.00351.083 0.41710.123↓↓↓↓0.0023
STAT11.097 0.48551.5230.04562.790↑↑0.0370
CDH10.993 0.78301.267 0.19726.590↑↑↑0.0001
CTNNB12.403↑↑0.00080.853 0.133410.96↑↑↑↑0.0001
FN13.673↑↑0.00390.843 0.14432.403↑↑0.0005
PDGFC0.6000.01103.247↑↑0.00162.813↑↑0.0033
PDGFRB3.350↑↑0.00284.733↑↑↑0.00023.010↑↑0.0069
VEGFC1.5830.02492.990↑↑0.00070.237↓↓↓0.0212
MAPK30.617 0.09472.313↑↑0.00064.903↑↑↑0.0044
SHC12.607↑↑0.00621.2500.003012.06↑↑↑↑0.0002
IRS215.74↑↑↑↑0.00014.457↑↑↑0.00014.300↑↑↑0.0003
TSC20.840 0.20962.453↑↑0.00344.810↑↑↑0.0012
MMP21.210 0.47242.573↑↑0.00571.2170.0097
TLR23.913↑↑0.00459.013↑↑↑↑0.00000.927 0.6464
MDM21.300 0.60301.427 0.19937.947↑↑↑0.0001
1 The genes are arranged according to their main functions or known molecular phenomena involved. 2 Averaged gene expression ratios (tumor v matching normal colon, n = 3). Significantly upregulated mRNA (arrows point up), downregulated mRNAs (arrows pointing down) (p < 0.05). Arrows emphasize the degree of differential expression (one arrow indicate <2 fold difference, two arrows indicate 2 to 4 fold difference, three arrows indicate 4 to 8 fold difference, four arrows indicate over 8 fold difference, all at p < 0.05).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kamel, F.; Schneider, N.; Nisar, P.; Soloviev, M. Bottom-Up Approach to the Discovery of Clinically Relevant Biomarker Genes: The Case of Colorectal Cancer. Cancers 2022, 14, 2654. https://doi.org/10.3390/cancers14112654

AMA Style

Kamel F, Schneider N, Nisar P, Soloviev M. Bottom-Up Approach to the Discovery of Clinically Relevant Biomarker Genes: The Case of Colorectal Cancer. Cancers. 2022; 14(11):2654. https://doi.org/10.3390/cancers14112654

Chicago/Turabian Style

Kamel, Faddy, Nathalie Schneider, Pasha Nisar, and Mikhail Soloviev. 2022. "Bottom-Up Approach to the Discovery of Clinically Relevant Biomarker Genes: The Case of Colorectal Cancer" Cancers 14, no. 11: 2654. https://doi.org/10.3390/cancers14112654

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