Next Article in Journal
Construction of Metastasis Prediction Models and Screening of Anti-Metastatic Drugs Based on Pan-Cancer Single-Cell EMT Features
Previous Article in Journal
Special Issue “Molecular Insights into the Developmental Origins of Health and Disease”
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Bioinformatic Approach to Identify Positive Prognostic TGFB2-Dependent and Negative Prognostic TGFB2-Independent Biomarkers for Breast Cancers

1
Oncotelic Therapeutics, 29397 Agoura Road, Suite 107, Agoura Hills, CA 91301, USA
2
Westmorland Campus, Kendal College, Market Place, Kendal, Cumbria LA9 4TN, UK
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2025, 26(23), 11580; https://doi.org/10.3390/ijms262311580 (registering DOI)
Submission received: 5 September 2025 / Revised: 11 November 2025 / Accepted: 26 November 2025 / Published: 29 November 2025
(This article belongs to the Section Molecular Oncology)

Abstract

Breast cancer is highly heterogeneous, with multiple subtypes that differ in molecular and clinical characteristics. It remains the most common cancer among women worldwide. We conducted a hypothesis-generating study using a bioinformatics approach in order to identify potential prognostic biomarkers for breast cancer patients across multiple molecular subtypes. Given the influential role of the transforming growth factor beta (TGFB) pathway in shaping the immune microenvironment, we focused on the isoform, transforming growth factor beta 2 (TGFB2), which is upregulated in tumors, to identify TGFB2-dependent and -independent biomarkers for breast cancer patients’ overall survival (OS) responses. We evaluated the impact of TGFB2 mRNA expression, in conjunction with other potential prognostic markers, on overall survival (OS) in breast cancer patients using The Cancer Genome Atlas (TCGA) and KMplotter databases. We employed a multivariate Cox proportional hazards model to compute hazard ratios (HRs) for TGFB2 mRNA expression, integrating an interaction term that accounts for the multiplicative relationship between TGFB2 and marker gene expressions while controlling age at diagnosis and cancer subtype and differentiating between patients receiving chemotherapy alone and those undergoing alternative therapeutic interventions. We used the KMplotter database to confirm TGFB2-independent prognostic markers from TCGA data. In cases dependent on TGFB2, increased mRNA expression of TGFB2 alongside higher levels of GDAP1, TBL1XR1, RNFT1, HACL1, SLC27A2, NLE1, or TXNDC16 was correlated with improved OS among breast cancer patients, of which four genes were upregulated in tumor tissues (SLC27A2, TXNDC16, TBL1XR1, GDAP1). Future studies will be required to confirm breast cancer patients could improve OS outcomes for patients expressing high levels of TGFB2 and the marker genes in prospective clinical trials. Additionally, multivariate analysis revealed that the elevated expression of six genes (ENO1, GLRX2, PLOD1, PRDX4, TAGLN2, TMED9) were correlated with increases in HR, independent of TGFB2 mRNA expression; all except GLRX2 were identified as druggable targets. Future investigations assessing protein expression in breast cancer tumors to confirm the results of our retrospective analysis of mRNA levels will determine whether the protein products of these genes represent viable therapeutic targets. Protein–protein interaction (STRING) analysis indicated that TGFB2 is associated with EGFR and MYC from the PAM50 breast cancer gene signature. These findings suggest that correlation of TGFB2-related markers could potentially complement the PAM50 signature in the assessment of OS prognosis in breast cancer patients, but further validation of the TGFB2/EGFR/MYC proteins in tumors is warranted.

1. Introduction

Breast cancer is characterized as highly heterogeneous, comprising various subtypes that exhibit distinct molecular and clinical features. The incidence of breast cancer has been notably high among women globally, making it the most common type of cancer in this demographic. Research has identified bimodal age distribution patterns at diagnosis, with peaks around ages 45 and 65, suggesting the existence of two main etiological subtypes influenced by shared risk factors [1]. Characterization of breast cancer subtypes has evolved using molecular subtyping techniques, notably the PAM50 classification, which categorizes breast cancers into intrinsic subtypes, including luminal A, luminal B, HER2+, and triple-negative breast cancer (TNBC). Each subtype is associated with differing prognoses and therapeutic responses. For example, the luminal A subtype is often linked to better outcomes, while TNBC typically exhibits a higher mutational burden and poorer prognosis [2,3,4]. Additionally, recent studies have highlighted the role of immune profiles and biological processes in further stratifying these subtypes. For instance, a high-risk luminal A subtype has been identified with increased motility and immune dysfunction, correlating with poor prognosis [2]. Furthermore, the Breast Cancer Consensus Subtypes (BCCS) have classified tumors based on gene expression, revealing that estrogen receptor-positive (ER+) subtypes like PCS1 and PCS4 are associated with better prognoses, whereas subtypes like PCS2 and PCS3 are linked to poor outcomes [3].
We focused on the orchestrating role of the family of signaling molecules known as transforming growth factor beta (TGFB) to identify potential overall survival (OS) prognostic biomarkers in cancer patients. TGFBs regulate cell differentiation, proliferation, and selectivity through various signaling pathways. The TGFB signaling pathway plays a complex, two-phase role in cancer development. In the early stages, TGFB exhibits suppressive effects by inducing apoptosis and inhibiting cell cycle progression, thereby reducing the proliferation of cancer cells. However, as the disease progresses, cancer cells develop resistance to TGFB signals. Instead of inhibiting tumor growth, TGFB initiates cancer-promoting processes, including invasion, immunosuppression, angiogenesis, and the formation of metastasis [5,6,7,8,9]. Our previous studies identified the specific role of the transforming growth factor beta 2 (TGFB2) mRNA isoform as a negative prognostic marker in pediatric brainstem diffuse midline glioma tumors. They showed that high TGFB2 mRNA levels were associated with an immunologically “cold” tumor microenvironment (TME), characterized by low expression of antigen-presenting cell (APC) markers (e.g., CD14, CD163, ITGAX/CD11c), which impairs anti-tumor immune responses [10]. In pancreatic ductal adenocarcinoma (PDAC), significant alterations in gene expression patterns were observed, particularly with a marked upregulation of TGFB2 mRNA, which demonstrates a 7.9-fold increase over normal pancreatic tissue [11]. Additionally, interferon pathway genes, such as IFNAR1, IRF9, STAT1, and IFI27, were significantly upregulated, with IFI27 exhibiting a highly significant 66.3-fold increase compared with normal tissue [11]. Notably, within TME characterized by low macrophage infiltration, high TGFB2 mRNA expression proves particularly detrimental to survival, highlighting a median overall survival of 15.3 months compared with 72.7 months for tumors with low TGFB2 expression [11]. In low-grade gliomas (LGGs), elevated mRNA levels of TGFB2 emerged as a significant negative prognostic marker, particularly when coupled with the activation of interferon-gamma receptor signaling via interferon regulatory factor 5 (IRF5) and the increased expression of the immune checkpoint molecule B7-H3 [12].
We have now extended these studies to examine the role of TGFB2 mRNA in breast cancer prognosis. Published studies have shown that the roles of transforming growth factor isoforms TGFB1, TGFB2, and TGFB3 in breast cancer prognosis are context-dependent [13,14,15,16,17,18]. Analysis revealed that both univariate (p = 0.01) and multivariate (p = 0.013) models showed serum TGFB1 levels that have a positive prognostic impact on overall survival in breast cancer patients [17]. Additionally, a study indicated that patients lacking TGFB1 expression experienced a 4.6 times higher risk of distant recurrence [16]. A comparative analysis of TGFB1 and TGFB3 immunoreactivity demonstrated co-expression of both isoforms in 111 tumors (72.5%), while 16 cases (10%) exhibited no expression. Immunostaining revealed that TGFB3, but not TGFB1, was inversely associated with overall survival (p = 0.0204). Furthermore, when considered in conjunction with lymph node involvement, TGFB3 emerged as a significantly more prognostic indicator for overall survival (p = 0.0003) [18]. The gene expression of TGFB peptides and receptors were analyzed in both normal and malignant breast tissue, including primary epithelial and stromal cultures. All TGFB isoforms and their receptors were expressed in normal breast samples, which expressed TGFB1 and TGFB3 (88% and 89%), while TGFB2 was less frequent (68%). Tumor stromal cells showed higher TGFB1 production than normal cells (p < 0.0001). In tumor-derived stromal cultures, TGFB1 and TGFB2 levels were strongly correlated (r = 0.976) [15]. Investigation of a total of breast cancer biopsy specimens using immunohistochemical analysis assessing expression levels of TGFB receptors TGFBR1 and TGFBR2 showed that TGFBR1 expression was detected in 273 out of 555 samples (49%), while TGFBR2 expression was observed in 239 out of 474 cases (50%) [19].
We utilized a bioinformatic-driven approach to characterize the impact of TGFB2 mRNA, in combination with potential prognostic markers, on overall survival in breast cancer patients from The Cancer Genome Atlas (TCGA) database. We implemented a multivariate Cox proportional hazards model to directly compare hazard ratio (HR) calculations for TGFB2 mRNA, an mRNA product of a marker gene (Gene2) expression, including an interaction term of TGFB2 by marker gene expression while controlling for age at diagnosis and breast cancer subtypes and comparing patients who received chemotherapy-only (chemo-only) therapies, and we filtered genes that reported HR greater than 1 for patients receiving only chemotherapy in the multivariate model. The interaction term in the model enabled the identification of TGFB2/Gene2 mRNA combinations that result in synergistic improvements for breast cancer patients. This resulted in TGFB2-dependent positive prognostic markers that could be potentially used as inclusion criteria in biomarker-guided clinical trial designs. We also identified TGFB2-independent Gene2, showing a negative correlation between mRNA expression and OS across all mRNA expression levels, and validated it with an independent dataset from the KMplotter database. The increased HR for patients with high levels of expression of these marker genes suggests that these marker genes could be presented as a future development for targeted therapies.

2. Results

2.1. TGFB2-Dependent Positively Prognostic Gene2

TGFB2 expression increased 1.95-fold in tumor tissues (Figure 1, p < 0.0001), indicating its involvement in tumor pathology. Using multivariate Cox analysis, we found that TGFB2 mRNA levels interact with Gene2 to affect the HR for overall survival (OS). This analysis was controlled for age at diagnosis, breast cancer subtype, treatment (chemotherapy-only (chemo-only) versus all other treatments), and the interaction between TGFB2 by Gene2 for 786 evaluable patients. Out of the 15,947 Gene2 screened, 380 genes exhibited significant interaction terms (p < 0.05), of which 111 genes exhibited significant upregulation in tumor tissues (Figure S1, p < 0.0001, FDR < 0.001).
In our first screen, we sought to identify TGFB2-dependent Gene2 effects that exhibited significant HR calculations for TGFB2 mRNA expression (p < 0.05) AND Gene2 expression (p < 0.05) AND TGFB2 by Gene2 interaction terms (p < 0.05) (Figure 1; 11 Gene2 upregulated in tumors). Expression of collagen type x alpha 1 chain (COL10A1) mRNA exhibited a significant (p < 0.0001) 877.9-fold change in tumor compared with normal tissue; this high value for fold change was a result of the very low expression of COL10A1 in normal tissue (mean ± SEM = −3.92 ± 0.16 log2TPM (transcripts per million)). Three genes showed a greater than 3-fold increase in expression in tumors compared with normal tissue: integrin subunit alpha 11 (ITGA11), sulfatase 1 (SULF1), and hydroxysteroid 17-beta dehydrogenase 6 (HSD17B6), all with p < 0.0001 (Figure 1B).
Examination of the multivariate Cox proportional hazards model revealed two sets of Gene2 markers from the first screen (Table S1). Six of these genes showed an increase in HR even when considering the positive prognostic impact of TGFB2 mRNA and significant interaction terms in the Cox model (Table S1): HSD17B6; ITGA11; glypican 4 (GPC4); COL10A1; ANTXR cell adhesion molecule 1 (ANTXR1); and SULF1 (HR ranged from 1.43 to 1.22). Five genes showed a decrease in the HR of both TGFB2 and Gene2 using the multivariate Cox proportional hazards model that is multiplied, as suggested by the significant TGFB2 by Gene2 statistical interaction effects: armadillo repeat containing 7 (ARMC7); transmembrane protein 14B (TMEM14B); autocrine motility factor receptor (AMFR); arylformamidase (AFMID); and ganglioside-induced differentiation-associated protein 1 (GDAP1), with the HR ranging from 0.02 to 0.62 (Table S1).
We next focused on Gene2 expression that exhibited positive prognostic impact in combination with TGFB2 mRNA expression from the list of 380 Gene2s that showed significant interaction parameters from the multivariate Cox proportional hazards models (Table S2). Seven Gene2s: GDAP1, TBL1X/Y related 1 (TBL1XR1), ring finger protein, transmembrane 1 (RNFT1), 2-hydroxyacyl-CoA lyase 1 (HACL1), solute carrier family 27 member 2 (SLC27A2), notchless homolog 1 (NLE1), and thioredoxin domain-containing 16 (TXNDC16) showed significant improvements in OS outcomes when comparing TGFB2high/Gene2high versus TGFB2high/Gene2high-expressing cohorts of patients using a Kaplan–Meier analysis; this was shown by separation for two arms of the four survival curves (p < 0.05), whereby high levels of TGFB2 and Gene2 mRNA expression resulted in the most improved OS outcomes (Figure 2, Table S3). Calculation of the multivariate parameters for these seven Gene2 reported p-values for the interaction term ranging from 0.004 to 0.036 (Table S2). Kaplan–Meier analysis of patient cohorts with above median expression of TGFB2 and GDAP1 mRNA (n = 181; p = 0.0084, Figure 2A); TBL1XR1 mRNA (n = 222; p = 0.0121, Figure 2B); RNFT1 mRNA (n = 190; p = 0.0164, Figure 2C); HACL1 mRNA (n = 184; p = 0.0188, Figure 2D); SLC27A2 mRNA (n = 177; p = 0.03, Figure 2E); NLE1 mRNA (n = 196; p = 0.031, Figure 2F); and TXNDC16 mRNA (n = 185; p = 0.036, Figure 2G) showed significantly improved survival outcomes than patients cohorts expressing high mRNA levels of TGFB2 and low levels of Gene2 (Figure 2, Table S3). Four of these genes were significantly upregulated in tumor tissues (p < 0.0001 for all comparisons) (Figure 2H): SLC27A2, TXNDC16, TBL1XR1, and GDAP1.
We investigated the correlation of TGFB2 with gene signature proteins using the network analyses provided in the protein–protein interaction networks from the STRING database (Figure 3, Figure 4 and Figure 5). We integrated data from two screens analyzing TGFB2-dependent Gene2 prognostic impacts (Figure 1 and Figure 2), showing TGFB2 was associated with the network of GPC4, SULF1, ANTXR1, ITGA11, and COL10A1 via linker proteins via second shell interactors: biglycan (BGN), collagen type III alpha 1 chain (COL3A1), collagen type I alpha 2 chain (COL1A2), collagen type V alpha 2 chain (COL5A2), and thrombospondin 2 (THBS2) (Figure 3). The PAM50 gene signature was associated with TGFB2 via MYC proto-oncogene, BHLH transcription factor (MYC), and epidermal growth factor receptor (EGFR) (Figure 4). TGFB2 was associated with the network of the OncotypeDX gene signature via TGFB1 associations with CD68 molecule (CD68), progesterone receptor (PGR), estrogen receptor 1 (ESR1), BCL2 apoptosis regulator (BCL2), and erb-B2 receptor tyrosine kinase 2 (ERBB2) (Figure 5).

2.2. TGFB2-Independent Gene2 Correlated with Increase in HR

We next performed multivariate analyses using the Cox proportional hazards model to assess the individual effects of TGFB2 and Gene2 marker mRNA expression levels, identifying Gene2 that exhibited an increase in HR, independent of TGFB2 mRNA expression. This analysis was controlled for age at diagnosis, breast cancer subtype, the confounding effect of treatment (chemo-only versus all other treatments), and the interaction between TGFB2 and Gene2. Out of the 15,947 Gene2 screened, 1286 exhibited increases in HR for Gene 2 (p < 0.05), of which 118 genes exhibited increases in HR for Gene2 not impacted by TGFB2 expression (p-value for TGFB2 > 0.1 AND p-value for TGFB2 by Gene 2 interaction > 0.1 AND chemo-only parameter > 1.5 standard deviation units). We compared the mRNA expression in tumor tissues of 118 Gene2 markers relative to normal tissues, resulting in 41 genes being significantly upregulated in tumor tissues for 786 evaluable patients (Figure S2, Table S4; using the filter: p < 0.0001, fold increase > 1, log2 (TPM) expression in tumor tissue > 2).
We cross-referenced the 41 genes identified in the TCGA dataset with those reported in the KMplotter database to narrow down the potential list of prognostic markers using more stringent criteria, comparing only median cut-off values for high- versus low-expressing patient subsets. This analysis identified six significantly upregulated genes in breast cancer tumor tissues cross-referenced in both the TCGA and KMplotter datasets (Figure 6). A significant increase in mRNA expression levels was observed in tumor tissue compared with normal tissue for the following genes: TAGLN2 (2-fold, p < 0.0001), ENO1 (1.49-fold, p < 0.0001), GLRX2 (1.73-fold, p < 0.0001), PLOD1 (1.72-fold, p < 0.0001), PRDX4 (1.81-fold, p < 0.0001), and TMED9 (2.28-fold, p < 0.0001) (Figure 6B).
To predict how Gene2 mRNA expression affects survival, we generated survival curves at any given time by plotting and calculating the shift in the baseline OS curve for 789 breast cancer patients (59 death events) by applying the parameters from the output of the multivariate regression model and then utilizing the fitted parameters (Table S5) to shift the baseline curves using values fixed for the control variables and varying a single variable of interest, thereby evaluating the Gene2 mRNA-dependent impact on OS. The shift in the baseline survival curves represented the median OS times throughout the entire range of Gene2 expression levels (minimum, lower quartile, median, upper quartile, and maximum), as well as the magnitude of the HR calculation. These simulated survival curves measured the impact of Gene2 alone, accounting for both the range of mRNA expression and the hazard ratio calculations (Figure 7, Table 1). Five genes exhibited HRs greater than 1.3 (30% increase in hazard for one standard deviation unit (Zscore) increase in mRNA expression): ENO1, GLRX2, PLOD1, PRDX4, TMED9, and TAGLN2 showed an HR of 1.7, controlling for TGFB2 mRNA expression, chemotherapy treatment, age at diagnosis, and subtypes of breast cancer (Figure 7, Table 1).
We next validated the TCGA prognostic markers, which was conducted using the independent KMplotter dataset for breast cancer patients with a larger sample size. The analysis focused on six genes (ENO1/ENO1L1, GLRX2, PLOD1, PRDX4, TAGLN2, and TMED9), which were found to significantly influence OS in both the TCGA and KMplotter datasets (Figure 8). At the median cut-off of mRNA expression for Gene2, high levels of expression showed worse OS outcomes for the six gene products identified in the TCGA and KMplotter datasets.

3. Discussion

In our first screen, the expression of 11 Gene2 markers exhibited significant prognostic impact for TGFB2 mRNA expression (p < 0.05) AND Gene2 expression (p < 0.05) AND TGFB2 by Gene2 interaction terms (p < 0.05). Six of these genes showed an increase in HR even when considering the positive prognostic impact of TGFB2 mRNA and significant interaction terms in the Cox model: HSD17B6; ITGA11; GPC4; COL10A1; ANTXR1; and SULF1. Notably, expression of COL10A1 mRNA exhibited a significant increase in tumor tissues, suggesting that expression of this gene is highly tumor-specific. The published literature has reported that in colorectal cancer, COL10A1 expression is associated with tumor progression and metastasis [20], and high expression of COL10A1 has been linked to poor overall survival and disease-free survival (DFS) in colorectal cancer patients [21]. In lung cancer, COL10A1 expression is associated with tumor progression and metastasis [22]. COL10A1 expression was found to be an independent risk factor for gastric cancer [23]. Interestingly, the circulating expression level of COL10A1 is significantly increased in gastric adenocarcinoma patients and associated with poor survival in these cancers [23]. Survival analysis of pancreatic cancer patients suggested that high expression of COL10A1 was associated with a poorer prognosis and that knockdown of COL10A1 inhibited the proliferation, migration, and invasion of cells in functional assays [24].
SULF1 mRNA exhibited a significant increase in breast tumor tissue. In our multivariate Cox proportional hazards model, increasing levels of SULF1 increased the HR in breast cancer patients (22% increase for one standard deviation increase in gene expression), in contrast to the positive prognostic impact of TGFB2 mRNA levels. Network analysis of SULF1 exhibited a second-order connection to TGFB2 via BGN, COL3A1, COL1A2, COL5A2, and THBS2 proteins.
SULF1 is a member of the sulfatase family, which regulates the sulfation of heparan sulfate proteoglycans and has been implicated in the prognosis of various types of cancer for both positive and negative prognostic impact. SULF1 mRNA expression was associated with a poor prognosis in lung adenocarcinoma [25]. A study that employed quantitative real-time polymerase chain reaction (qRT-PCR) to measure SULF1 mRNA expression from 54 non-small-cell lung cancer (NSCLC) patients showed that patients expressing high levels of SULF1 mRNA exhibited shorter median OS times. Furthermore, knockdown of SULF1 suppressed key malignant behaviors in NSCLC cell lines, including NCI-H1299 (CRL-5803) and HCC827 (CRL-2868), by reducing cell proliferation, migration, invasion, and epithelial–mesenchymal transition (EMT) through inhibition of the EGFR/MAPK signaling pathway [26]. In hepatocellular carcinoma (HCC), SULF1 promotes TGFB-induced gene expression and EMT transition [27]. In vitro studies demonstrated that forced SULF1 expression in HCC cell lines (Hep3B, PLC/PRF/5) resulted in increased SMAD2/3 phosphorylation following stimulation with TGFB1. Conversely, SULF1 knockdown in cell lines with high endogenous levels (SNU182, SNU475) reduced TGFB signaling, migration, and invasiveness [27]. Furthermore, both immunohistochemistry (IHC) and Western blotting identified elevated SULF1 expression and increased phosphorylation of SMAD2/3 (TGFB pathway activation) in tumor and peritumoral tissues from transgenic Sulf1-Tg mice compared with WT [27]. SULF1 expression facilitated the upregulation of the hallmarks of EMT mesenchymal markers: N-cadherin, vimentin, and αSMA [27]. Notably, these observations were based on the activation of TGFB1 expression, whereas our prognostic model suggested that the TGFB2 mRNA isomer drove a positive prognostic impact in breast cancers, indicating the need for a careful and nuanced examination of the TGFB ligand isoforms in tumor progression. In gastric cancer, [28] Fang et al., 2024, reported that SULF1, produced by cancer-associated fibroblasts (CAFs), facilitated metastasis and resistance to cisplatin treatment via TGFB1 activation of TGFBR3-mediated signaling. Pan-cancer analysis revealed that SULF1 mRNA was upregulated more than 2-fold in 16 out of the 32 cancers in the TCGA dataset [29]. Analysis of a single-cell RNA-seq (scRNA-seq) dataset of head and neck squamous cell carcinoma (HNSC) showed the highest positivity of SULF1 in fibroblasts [29]. In pancreatic cancer, high SULF1 expression has been associated with later T, N, and TNM stages, higher CA19-9 levels, smaller tumor size, and poorer prognosis [30].
There is a tumor suppressor role for SULF1, being an extracellular sulfatase that removes 6-O-sulfate groups from heparan sulfate (HS) chains, thereby reducing the activity of HS-binding growth factors such as FGF2, VEGF, amphiregulin, HB-EGF, and HGF [31]. By removing these sulfate groups, SULF1 decreases growth factor presence at the cell surface and restricts receptor interaction, ultimately suppressing pro-tumorigenic signaling [31].
In breast cancers under hypoxic conditions, SULF1 is downregulated by HIF-1α, promoting cancer cell migration and invasion, in aggressive breast cancer cell lines [32]. In this study, a Kaplan–Meier survival analysis showed that tumors with high SULF1 mRNA expression exhibited longer DFS compared with patients whose tumors had low levels of SULF1 mRNA expression [32]. Ovarian cancer cell lines and primary tumors lacking SULF1 mRNA expression showed dense DNA methylation in 12 CpG sites within exon 1A and increased histone H3 methylation around the SULF1 gene promoter, leading to transcriptional repression [33]. Downregulation of SULF1 mRNA in ovarian cancer cells, achieved via siRNA, reduces sensitivity to cisplatin-induced cytotoxicity, indicating that loss of SULF1 mRNA promotes chemoresistance. Furthermore, loss of SULF1 leads to decreased expression of the pro-apoptotic protein Bim via increased ERK signaling, promoting tumor cell survival and resistance to chemotherapy [34].
Taken together, the impact of SULF1 in the TME is highly context-dependent on the tumor type, expression in stromal cells, the mechanism of SULF1 gene activation, and hypoxia. SULF1 interacts with the TGFB pathway via TGF-beta 1. Our analysis’s high levels of SULF1 are overexpressed in tumors and have a negative prognostic impact in the context of low levels of a different TGFB isomer, TGFB2, suggested by the significant statistical interaction term calculated from the multivariate Cox proportional hazards term. The network analysis showed no direct mechanistic connection of TGFB2 and SULF1, but they are components of a highly interconnected set of associations. Further work will need to elucidate the expression of TGFB1, TGFB2, and SULF1 using scRNAseq studies for the distribution of expression in the cellular compartments in tumors.
Expression of ITGA11 mRNA exhibited a significant (p < 0.0001) 3.88-fold increase in tumor tissues. ITGA11 mRNA expression has been found to play a significant role in cancer prognosis. Studies have shown that high expression of ITGA11 is associated with poor prognosis in various types of cancer, including gastric cancer [35], NSCLC [36], and breast cancer [37]. In breast cancer, ITGA11 has been found to play a role in the regulation of cancer-associated fibroblasts, tumor progression, and prognosis [37,38].
Five genes showed positive prognostic impact of both TGFB2 and Gene2 in that combination is enhanced as suggested by the significant TGFB2 by Gene2 interaction effects: ARMC7; TMEM14B; AMFR; AFMID; and GDAP1. Our analysis revealed a novel positive TGFB2-dependent effect of these Gene2 markers on breast cancer survival, whereas previous studies on the expression of these TGFB2-dependent genes focused solely on their impact on either the risk score or as negative prognostic biomarkers.
ARMC7 is one of the seven differentially expressed mRNAs included in a prognostic signature for predicting recurrence and DFS in cervical cancer patients [39]. High plasma levels of a five-gene signature, which included ARMC7, were significantly associated with inferior overall survival in diffuse large B-cell lymphoma patients, independent of the National Comprehensive Cancer Network International Prognostic Index score [40].
AMFR has been shown to have significant prognostic implications in various cancers, primarily indicating a poor survival outcome in NSCLC [41], gastric cancer [42], and a borderline statistically significant correlation with shorter overall survival (p = 0.0331) in breast cancers [43]. Interestingly, in colorectal cancer AMFR overexpression correlated with improved DFS in colon cancer but was associated with decreased DFS in corresponding nodal metastases. In rectal cancer, AMFR overexpression was linked to significantly reduced OS, DFS, and disease-specific survival (p < 0.001, p = 0.031, p = 0.005, respectively) [44]. This supports our observation of the context-dependent impact of AFMR expression on clinical outcomes in cancer patients.
In our second screen, we identified TGFB2-dependent positive prognostic markers from the survival dataset directly by performing a Kaplan–Meier analysis using median cut-offs for TGFB2 and the paired Gene2 expression (GDAP1, TBL1XR1, RNFT1, HACL1, SLC27A2, NLE1, and TXNDC16). Four of these genes’ mRNA were significantly upregulated in tumor tissues (p < 0.0001 for all comparisons) (Table S4; SLC27A2 (6.78-fold increase); TXNDC16 (1.85-fold increase); TBL1XR1 (1.81-fold increase), and GDAP1 (1.53-fold increase)).
GDAP1 is primarily involved in mitochondrial function and is localized to the outer mitochondrial membrane [45]. A high level of GDAP1 expression was correlated with improved survival in pancreatic cancer patients (HR: 0.71 [95% CI: 0.57–0.90]; p = 0.004) [46]. This contrasts with the negative prognostic marker in acute myeloid leukemia [47]. GDAP1 contributed to a gene signature panel correlated with risk prediction in hepatitis B virus-associated HCC [48]. Our studies indicated that the prognostic impact was TGFB2 mRNA-dependent using both the Cox proportional hazards model and the Kaplan–Meier analysis, suggesting that breast cancer patients with high levels of both TGFB2 and GDAP1 expression displayed significantly improved OS outcomes. Therefore, patients expressing high mRNA levels of these two genes displayed improved OS outcomes to standard therapies, which requires confirmation in prospective clinical trials.
SLC27A2 is primarily involved in the cellular uptake and activation of long-chain and very long-chain fatty acids in human tissues [49]. In breast cancers, SLC27A2 mRNA expression constituted part of a four-gene signature that stratified patients into low and high-risk groups [50]. Increased levels of SLC27A2 were validated at both the mRNA and protein levels in breast cancer cell lines, and the knockdown of SLC27A2 inhibited cell proliferation and cell cycle in these cells [50]. The oncogenic role of SLC27A2 was demonstrated in differentiated thyroid cancer [51] and acute lymphoblastic leukemia [52]. Expression of SLC27A2 was found to be context-dependent based on its tumor-suppressive role in renal cell carcinoma [53]. In this cancer, SLC27A2 influences EMT transition by negatively regulating CDK3, whose low expression is associated with a poor prognosis [53]. Our studies also suggested that the prognostic effect of SLC27A2 was context-dependent on TGFB2 mRNA levels, whereby high levels of SLC27A2 and TGFB2 mRNA levels showed improved OS outcomes.
TXNDC16 is a thioredoxin domain-containing protein that belongs to the larger thioredoxin superfamily, which plays a role in supporting cellular redox homeostasis, proper protein folding, and electron transfer. Members of the thioredoxin superfamily of proteins have been reported to be influential in glioma pathogenesis [54]. TBL1XR1 is a multifunctional protein, and its functions are cell type- and context-dependent. It is understood to be involved in transcription and a range of signaling pathways [55]. A recent review identified 12 different cancers, including breast cancer, which implicate TBL1XR1 in cancer progression [56]. In breast cancer, TBL1XR1 was correlated with clinical stage, tumor classification, node classification, metastasis classification, and histological grade, and higher TBL1XR1 expression was an independent prognostic indicator for overall survival. The mechanism may be via cyclin D1-transactivation and activation of the β-catenin signaling pathway [55].
Three genes (NLE1, HACL, and RNFT1) did not exhibit significant upregulation in tumor tissues. In contrast to the positive prognostic impact of these genes in combination with TGFB2, NLE1 is elevated and positively correlated with tumor stage and metastasis in melanoma [57]. Up-regulation of the NLE1 mRNA in colon cancer cells promotes cell migration and invasion of colon cancer cells and inhibits the apoptosis ability of colon cancer cells [58]. NLE1 has also been implicated in the progression of non-small-cell lung cancer. NLE1 expression is significantly higher in tumor tissues than in normal tissues, is correlated with the pathological stage, and can be considered a tumor promoter [59]. Finally, in HCC, high NLE1 expression has been shown to promote cell proliferation and to be strongly associated with poor prognosis [60].
Investigation of the TGFB2-dependent gene signatures for their protein–protein associations from the two screens in our study demonstrated that TGFB2 was associated with a network consisting of GPC4, SULF1, ANTXR1, ITGA11, and COL10A1 via the linker proteins BGN, COL3A1, COL1A2, COL5A2, and THBS2. Studies have demonstrated that TGFB induces BGN expression in pancreatic cells through the activation of MKK6-p38 MAPK signaling downstream of Smad signaling, offering an insight into the regulation of BGN observed in fibrosis and desmoplasia associated with inflammatory responses [61]. Furthermore, the TGF-beta/ALK5 effect on p38 activation and BGN expression was also impacted by overexpression of GADD45beta alone in PANC-1 and osteosarcoma MG-63 cells [62]. Regarding research directed towards identifying direct Smad targets in dermal fibroblasts following TGFB stimulation, requiring demonstration of a rapid increase in mRNA levels, TGFB-induced promoter activity blocked by dominant-negative Smad3 and Smad7 vectors, and no promoter transactivation in Smad3(-/-) fibroblasts found COL3A1 as a TGFB/Smad3 target [63]. TGFB2 treatment elevates COL1A2 expression in human epithelial SRA01/04 cells, whereby COL1A2 knockdown inhibited TGFB2-induced SRA01/04 cell proliferation, migration, invasion, and EMT [64]. COL5A2 expression was shown to be elevated in human osteosarcoma cells, and the downregulation of COL5A2 affected the TGFB and Wnt/β-catenin signaling pathways [65]. THBS2 expression in pancreatic cancer is mainly present in the stroma and is linked to tumor progression and poor prognosis. RNA in situ hybridization revealed that cancer-associated fibroblasts (CAFs), not tumor cells, express THBS2, with levels increasing as the disease progresses in mouse models [65]. TGFB1 from cancer cells activated CAFs to produce THBS2 through the p-Smad2/3 pathway, suggesting an important role for TGFB association with THBS2 in cancer progression [66]. It is noted that the TGFB2 association represents a strong correlation; future controlled experiments in reduced systems will be required to infer mechanistic insight of this TGFB2 association in breast cancer tissues.
Since our multivariate prognostic model identified TGFB2-dependent prognostic markers across the TCGA cohorts including unclassified, basal, normal, luminal A, luminal B, and HER2+, we performed a STRING network analysis on the PAM50 gene signature that classifies cancers according to luminal A, luminal B, HER2+, basal-like, and normal-like, and it has been further developed by combining risk scores to improve future recurrence risk [67,68]. Our analysis revealed that TGFB2 was linked to the network of the PAM50 gene signature through MYC and EGFR. In cancer cell lines, studies have demonstrated that genetic or pharmacological inhibition of MYC in MCF10A basal breast cells results in increased sensitivity to TGFB-stimulated invasion and metastasis [69], and signaling via Smad3 and ERK/Sp1 mediate TGFB-induced EGFR upregulation [70]. In particular, from our studies, the mRNA expression of GPC4, SULF1, ANTXR1, ITGA11, and COL10A1 exhibited significant statistical interactions with TGFB2 mRNA. TGFB2 was a positive prognostic factor, and the five genes were negative prognostic factors from the HR calculations of multivariate models (Table S1). Our results suggest that combinations of high levels of TGFB2 mRNA and low levels of these five genes resulted in improving patient OS outcomes. This correlation suggests that TGFB2 mRNA expression levels in breast cancer tumors may be used in conjunction with the PAM50 gene signature to assess the prognostic impact on patients following validation in prospective clinical trials.
OncotypeDx is a genomic test used for early-stage, hormone receptor-positive, HER2-negative breast cancer to help guide treatment decisions, specifically regarding chemotherapy. It analyzes the expression of 21 genes in a tumor sample to provide a recurrence score from 0 to 100. This score indicates the risk of cancer returning and the potential benefit from chemotherapy [71]. The STRING network analysis showed associations of this gene signature with TGFB2 via TGFB1 associations with CD68, PGR, ESR1, BCL2, and ERBB2. TGFB1, ERBB2, and CD68 form a signaling network that impacts cancer progression, invasion, and resistance to therapy in HER2-overexpressing breast cancer tumors [72,73]. There was no direct association of TGFB2 mRNA and the OncotypeDX gene signature.
Expression of six genes (ENO1, GLRX2, PLOD1, PRDX4, TAGLN2, and TMED9) was identified to be significantly upregulated in tumor tissues, exhibiting an increase in HR on breast cancer patients analyzed in the TCGA dataset using the multivariate Cox proportional hazards model and validated in an independent KMplotter dataset.
In breast cancer, ENO1’s role can be context-dependent: in early-stage breast cancer, higher ENO1 expression correlates with better survival and increased anti-tumor immune infiltration, while in advanced disease, it is associated with progression and immune evasion [74]. ENO1 is frequently overexpressed and is consistently linked to poor prognosis, advanced stage, and aggressive tumor features across multiple malignancies, including breast, lung, bladder, prostate, gastric, colorectal, pancreatic, and thyroid cancers [75,76,77,78]. Mechanistically, ENO1 drives tumorigenesis by promoting glycolysis (the Warburg effect); activating oncogenic signaling pathways such as PI3K/Akt, β-catenin, TGFB/Smad, and Wnt; and facilitating epithelial–mesenchymal transition, angiogenesis, and immune evasion [79,80,81,82,83,84]. ENO1 also regulates apoptosis, cell cycle progression, and interacts with c-Myc [85]. ENO1’s cell surface localization and cancer-specific overexpression make it a promising and accessible biomarker for diagnosis and prognosis, as well as a therapeutic target. Preclinical studies demonstrate that targeting ENO1 using antibodies, small molecule inhibitors, DNA vaccines, or gene knockdown can suppress tumor growth, invasion, metastasis, and chemoresistance and may enhance the efficacy of chemotherapy [78,79,84,86,87]. Additionally, ENO1 influences the TME by modulating immune cell infiltration and promoting CD8+ T cell exhaustion in some cancers [76,88]. ENO1 is a central driver of cancer metabolism, progression, and immune modulation, making it a promising candidate for integrated diagnostic, prognostic, and therapeutic strategies in oncology [74,77,78]. Availability of a small, cell-permeable molecule inhibitor of ENO1 such as ENOblock/AP-III-a4 [89] suggests that breast cancer patients with high levels of ENO1 mRNA can be potentially targeted for treatment. ENO1 inhibition enhances anti-tumor immunity and sensitizes tumors to immune checkpoint blockade therapies, such as anti-PD-L1, in bladder cancer models. CRISPR-Cas9-mediated knockout of ENO1 and ENOblock reduced tumor burden and promoted infiltration of cytotoxic CD8+ T cells [90]. A citrullinated peptide vaccine targeting ENO1 (citENO1 vaccine) in triple-negative breast cancer (TNBC) mouse models showed synergistic effects with PD-1 blockade, significantly inhibiting tumor growth and improving survival compared with controls [91]. Our results suggest that ENO1 is targetable for treated patients with high levels of ENO1 mRNA. The prognostic impact at the mRNA level will need to be confirmed by measuring protein levels in the tumors.
GLRX2 maintains mitochondrial redox balance by modulating glutathionylation and interacting with thioredoxin/glutathione systems, protecting against oxidative stress-induced apoptosis [92]. It is implicated in a variety of other cancers. High GLRX2 expression correlates with better prognosis and survival in early-stage colon adenocarcinoma (87% in stage I vs. 1% in stage III), acting as an independent prognostic marker [93]. GLRX2 is upregulated in bladder cancer and contributes to diagnostic models, suggesting a pro-tumorigenic role [94]. In gliomas, high GLRX expression is associated with aggressive subtypes (wild-type IDH, methylated MGMT) and is involved in immune modulation [95]. Serum GLRX2 levels in colon cancer patients correlate with survival [93]. GLRX2 influences tumor-associated immune processes, particularly via M0 macrophages and immune checkpoint pathways in gliomas, indicating a role in shaping the TME [95]. Currently, no specific inhibitors to GLXR2 have been described in the literature. Our results suggest that breast cancer patients with low levels of GLXR2 expression exhibited improved OS outcomes with standard therapies. Prospective clinical trials examining GLXR2 mRNA levels will be required to validate our observations.
PLOD1 hydroxylates lysine residues in collagen, enabling cross-linking that strengthens connective tissues (e.g., skin, bones). PLOD1 is upregulated in osteosarcoma, gastric, breast, glioma, bladder, pancreatic, and hepatocellular cancers (HCCs), correlating with advanced tumor stages, metastasis, and reduced survival [96,97,98,99]. High PLOD1 expression predicts poor overall and DFS in multiple cancers, including breast cancer [100,101,102]), and PLOD1 inhibition (via siRNA or small molecules) attenuates proliferation, migration, and invasion in bladder cancer and glioma models [97,103]. PLOD1 drives proliferation, migration, invasion, and therapy resistance in cancers by modulating pathways like β-catenin, NF-κB, and Akt/mTOR [96,97,98]. In HCC, PLOD1 stabilizes SEPT2 via hydroxylation, promoting F-actin networks that enhance confined cell migration [99]. PLOD1 expression correlates with the infiltration of immunosuppressive cells (e.g., macrophages, fibroblasts) and reduced CD8+ T cell activity. It is also associated with immune checkpoints such as PD-L1 [100,104]. PLOD1 protein contains a prolyl 4-hydroxylase alpha subunit homologs (P4Hc) domain that can be targeted for inhibition by compounds akin to FG-4592 (roxadustat), GSK1278863 (daprodustat), molidustat, IOX4, and vadadustat [105]. Our study suggests that patients expressing high levels of PLOD1 can be targeted for therapies against prolyl 4-hydroxylase alpha subunits pending confirmation that an increase in PLOD1 mRNA results in increases in PLOD1 protein levels in breast cancer patients.
PRDX4 is a key endoplasmic reticulum (ER)-resident antioxidant enzyme involved in hydrogen peroxide metabolism, redox homeostasis, and oxidative protein folding, thereby protecting cells from oxidative and ER stress and supporting normal physiological processes such as protein folding and ovarian function [106,107,108,109]. In cancers, PRDX4 is frequently overexpressed in multiple tumor types, including breast, prostate, ovarian, colorectal, lung, and gastric cancers, and is associated with tumorigenesis, therapeutic resistance, metastasis, and recurrence [106,107,108,109,110]. Mechanistically, PRDX4 promotes cancer cell proliferation, migration, invasion, and radio resistance, often through pathways such as Akt/GSK3 and by modulating the TME [107,109,110]. Knockdown or depletion of PRDX4 in cancer cells leads to reduced proliferation, increased apoptosis, impaired migration and invasion, and heightened sensitivity to radiation and chemotherapy [107,109]. PRDX4 also influences the immune landscape of tumors. Its overexpression is linked to reduced lymphocyte infiltration and increased immunosuppression, partly through interactions with proteins like TXNDC5 and modulation of immune checkpoints [108]. In vivo studies in a silicosis mouse model demonstrated that pharmacological inhibition of PRDX4 using conoidin A (Con A) significantly improved lung function, reduced fibrosis and inflammatory infiltration, and decreased collagen deposition and lung damage [111]. Our correlative results suggest that breast cancer patients with high levels of PRDX4 may be targeted using compounds such as Con A, pending confirmation that increased protein levels are also prognostic in breast cancer patients.
TAGLN2 is essential for immune cell function, particularly in T cells and macrophages, where it stabilizes the actin cytoskeleton at the immunological synapse, enhances T cell activation, and promotes phagocytosis [112,113]. TAGLN2’s role in breast cancer metastasis was examined through transwell migration, luciferase, and flow cytometry assays, as well as a mouse xenograft model, suggesting TAGLN2 acts as a tumor suppressor, and its loss may enhance breast cancer metastasis via activation of the ROS/NF-κB pathway [114]. However, our results suggest that upregulation of TAGLN2 mRNA was correlated with worse prognosis in breast cancer patients and concurred with other studies, including gastric, colorectal, papillary thyroid, and endometrial cancers, whereby TAGLN2 is overexpressed and correlated with poor prognosis, advanced tumor stages, metastasis, and drug resistance [115,116,117,118]. It promotes proliferation, migration, invasion, angiogenesis, and EMT [117,118,119]. In breast cancer patients, the impact of TAGLN2 may operate via activation of the NRP1/VEGFR2 and MAPK pathways, as in gastric cancer, promoting angiogenesis and cell survival [119]; the Rap1/PI3K/AKT pathway, as reported in papillary thyroid cancer, to enhance invasion [117]; or, akin to colorectal cancer progression, by activating STAT3 and regulating ANXA2 [118]. A more careful examination of the signaling mechanisms from breast cancer patients will need to be elucidated to clarify the roles of tumor-suppressing and tumor-promoting signaling in these patients. Knocking down TAGLN2 may present as a target in breast cancer patients, based on reports that TAGLN2 knockdown increases DNA damage and sensitizes gastric cancer cells to chemotherapy and radiation [116]. Depleting TAGLN2 inhibits migration and invasion in endometrial cancer, and recombinant TAGLN2-based therapies reduce tumor growth in vivo, suggesting therapeutic potential [115]. Small molecule inhibitors that block transgelin-2-dependent actin polymerization have been found to significantly reduce growth and invasion of GBM 08-387 and 3359 PDX cell lines in vitro [120]. Although no clinical trials are testing these compounds, inhibition of TAGLN2 may be pursued as a therapeutic option for breast cancer patients expressing high levels of this gene following confirmation of the prognostic impact of the protein in breast cancer patients.
TMED9 is upregulated in breast cancer, HCC, ovarian cancer, and gliomas, correlating with advanced tumor stages, metastasis, and poor survival [121,122,123,124,125]. TMED9 drives colon cancer metastasis via the CNIH4/TGFα/GLI pathway, inducing EMT and enhancing migration/invasion [125,126]. TMED9 promotes oncogenic signaling via TGFα/GLI, EMT, and STAT3 pathways [125,126]. In gliomas, TMED9 correlates with immune cell infiltration (e.g., B cells, T cells) and immune checkpoints like CD274 (PD-L1), suggesting a role in immune evasion [127]. BRD4780 is a small molecule that targets TMED9 [128]. Our results suggest that breast cancer patients with high levels of TMED mRNA are potentially susceptible to inhibiting TMED pending confirmation of protein levels.
Applying bioinformatics, our hypotheses-generating study identified biomarkers correlated with OS in breast cancer patients, underscoring the importance of further laboratory validation of the marker mRNA and protein expression of these markers. Our observations suggest that biomarkers predictive at elevated TGFB2 mRNA levels may be influenced by their biological and biochemical context, warranting additional studies to clarify these relationships. Protein–protein interaction network maps revealed that TGFB2 correlates with EGFR and MYC from the PAM50 gene signature, indicating that protein-level validation of these genes is needed to determine if TGFB2-related markers can complement the PAM50 signature in future prospective clinical trials. Further validation will require cell-level expression in the TME to determine the localization of TGFB2 and the biomarkers, thereby elucidating the establishment of TGFB2 dependency correlated with OS outcomes. Thus, a validation pipeline combining qRT-PCR, IHC, and scRNA-seq studies will enable comprehensive confirmation of biomarkers by integrating bulk quantitative mRNA measurement, spatial protein localization, and single-cell resolution of tumor heterogeneity. Initial qRT-PCR on bulk tumor RNA quantifies marker mRNA levels and validates differential expression, followed by IHC, to analyze protein expression patterns and distribution in formalin-fixed tissues. Concurrently, scRNA-seq of fresh tumor samples will reveal cell type-specific expression, heterogeneity, and involvement in TME dynamics. Integrating these complementary approaches ensures robust validation by confirming molecular signatures predicted by bioinformatics, correlating mRNA and protein data, and elucidating the roles of biomarkers in distinct cellular subpopulations, thereby accelerating clinical translation and personalized therapy development in breast cancers.

4. Materials and Methods

4.1. AI-Powered Chatbot Identified PubMed Abstracts for Manuscript Preparation

PubMed searches using the keywords: “(ARMC7 OR TMEM14B OR AMFR OR AFMID OR GDAP1) AND cancer” (171 abstracts); “(GDAP1 OR SULF1 OR ITGA11 OR HSD17B6 OR COL10A1) AND cancer” (410 abstracts); “TGF-beta AND Breast” (3667 abstracts); and “Gene AND TCGA AND prognosis AND breast” (2273 abstracts) (https://pubmed.ncbi.nlm.nih.gov/, accessed on 18 December 2024) were downloaded as text documents for processing using the Oncotelic Chatbot technologies. We also utilized Perplexity AI chatbot to identify articles (https://www.perplexity.ai/, accessed on 27 October 2025) not deposited in PubMed. Detailed methods are described in the Supplementary Methods. Chatbot technology enabled interactive keyword refinement for PubMed keyword searches by conducting question-and-answer sessions to refine the references; however, it was not utilized for systematic review, result validation, hypothesis formulation, or manuscript writing. All biomarker discoveries in the manuscript resulted from bioinformatic analyses, and identification of these genes were used to interrogate PubMed and Perplexity AI tools to contextualize the literature for these genes.

4.2. Differential Expression of mRNA Comparing Breast Cancer Tumors Versus Normal Samples

We analyzed gene expression by downloading log2-transformed TPM RNA-seq summary files (see https://toil-xena-hub.s3.us-east-1.amazonaws.com/download/TcgaTargetGtex_rsem_gene_tpm.gz (accessed on 10 November 2025) for full metadata) from the UCSC Xena web platform (https://xenabrowser.net/datapages/, accessed on 25 July 2023) [129]. This enabled a comparative study of 179 normal breast tissue samples (“GTEX Breast”) and 786 breast cancer patient samples with associated clinical data (“TCGA Breast Invasive Carcinoma”). The data are sourced from the UCSC Toil RNA-seq recompute compendium, which provides a harmonized dataset with realigned and recalculated gene and transcript expression levels for all TCGA, TARGET, and GTEx samples [130], facilitating direct comparisons of gene expression between TCGA tumor samples and matched GTEx normal tissues.
The gene expression measure in the Xena database, especially for TCGA cohorts, is normalized, log2-transformed RSEM values represented as log2(x + 1), where x is the normalized expression value estimated by the RNA-Seq Expectation-Maximization (RSEM) method. This metric is designed to facilitate cross-sample comparisons by stabilizing variance and normalizing variations in sequencing depth. We applied ANOVA methods to determine differential gene expression values such as RSEM normalization and transformations for the following features: 1. variance stabilization to make the data more normally distributed, satisfying key ANOVA assumptions; 2. variance modeling by quantifying uncertainty and variance in expression estimates by incorporating read mapping ambiguity and fragment length distributions during expectation maximization; and 3. handling heteroscedasticity by modelling the variance mean dependency explicitly. Log-transforming RSEM values reduces heteroscedasticity (unequal variance), thereby improving the validity of ANOVA or related variance component tests on expression data. There are related methods (DESeq2/edgeR) that leverage count reads directly for gene-level comparisons. The Xena platform provides batch-corrected RSEM-normalized and -transformed expression values from two different datasets, including GTEx normal samples, TCGA tumor samples, and adjacent normal samples, which enables robust comparisons between normal and tumor samples.
We applied a two-way ANOVA model to identify differentially expressed genes to compare normal versus tumor tissue samples. The log2-transformed TPM values for gene and tissue were included as fixed factors, along with one interaction term to investigate gene-level effects for normal and tumor tissues (gene × tissue). For each gene, we conducted a comparison between normal and tumor samples blocked by the gene factor and then determined significance by adjusting the p-value using the false discovery rate (FDR) algorithm provided for in the R-package (FDR corrected for all pairs) calculations performed in R using multcomp_1.4-17 and emmeans_1.7.0 packages ran in R version 4.1.2 (1 November 2021) with RStudio front end (RStudio 2021.09.0+351 “Ghost Orchid” Release). Bar chart graphics were constructed using the ggplot2_3.3.5 R package.
We applied a two-way hierarchical clustering approach to arrange gene expression profiles, grouping together both samples and genes with comparable mRNA expression patterns. This was achieved by employing the average linkage method and the default Euclidean distance metric, as implemented via the heatmap.2 function from the R package gplots_3.1.1. The resulting cluster visualization illustrated the mean expression in tumor samples, normalized to the average expression observed in normal breast cancer tissue, and presented as log2-transformed fold-change values. Dendrograms were generated for both the rows (genes) and the columns (samples), thereby displaying the organization and relationship of co-expressed genes across the patient cohort.

4.3. OS Outcomes for Breast Cancer

We analyzed clinical metadata (https://www.cbioportal.org/study/summary?id=brca_tcga_pan_can_atlas_2018; accessed on 1 February 2024) and RNA sequencing data (“data_mrna_seq_v2_rsem.txt”: Batch normalized from Illumina HiSeq_RNASeqV2 using the RSEM algorithm (Transcripts per million (TPM)) from 789 patients diagnosed with breast cancer to compare OS correlates with mRNA expression values.
Multivariate analyses utilized the Cox proportional hazards model to assess the individual effects of TGFB2 and Gene2 mRNA expression levels (screened 15,947 genes) on overall survival (OS) to calculate the HR. This analysis was controlled for age at breast cancer subtype, diagnosis, treatment (chemotherapy-only versus all other treatments), and the interaction between TGFB2 and Gene2 mRNA. The inclusion of a statistical interaction term was an important consideration in our analysis design. Gene expression of TGFB2 mRNA and Gene2 mRNA were the main effects of the model. The interaction coefficients from the interaction term in the model represents the additional effect of the interaction term at a specific expression level of TGFB2 mRNA on Gene2 expression levels; the interaction term determines how much more the effect of TGFB2 mRNA levels impacts the effect of Gene2 mRNA expression. The HR associated with an interaction term is a multiplicative effect on the baseline hazard, meaning it represents a multiplicative change in the HR when both interacting conditions are present, compared with having only the main effects at defined mRNA levels for TGFB2 and Gene2. Briefly, the model included (i) TGFB2 mRNA levels as a linear covariate expressed as Zscores (n = 789); (ii) the mRNA expression level for Gene2 mRNA levels as a linear covariate expressed as Zscores (n = 789); (iii) TGFB2 by the Gene2 interaction term to determine the dependency of Gene2 prognostic impact on TGFB2 mRNA levels (n = 789); (iv) treatment comparing chemo-only versus other treatment regimens (n = 101 for chemo-only; n = 688 for other); (v) breast cancer subtype (unclassified as reference (n = 78), basal, normal (n = 24), luminal A (n = 371), luminal B (n = 143), and HER2+ (n = 50)); and (vi) age at diagnosis expressed as a linear covariate (n = 789), implemented in R (survival_3.2-13 ran in R version 4.1.2. Forest plots were utilized to visualize the hazard ratios for Cox proportional hazards models for OS outcomes (survminer_0.4.9 ran in R version 4.1.2 (1 November 2021). The breast cancer subtype factor in the model controlled for differential survival of the patients, thereby identifying gene expression markers with prognostic impact across all the subtypes reported in the database. The life table HRs were estimated using the exponentiated regression coefficient for Cox proportional hazards analyses implemented in R (survival_3.2-13 ran in R version 4.1.2).
To identify potential prognostic Gene2 mRNA markers, we considered only expression of Gene2 that exhibited HR values greater than 1 for chemo-only treated patients. We applied the multivariate Cox proportional hazards model to mRNA expression values across 15,947 genes and considered that Gene2 prognostic impact was dependent on TGFB2 mRNA levels with the statistical TGFB2 by the Gene2 interaction parameter p-value being less than 0.05. To better understand different combinations of TGFB2 mRNA levels and Gene2 mRNA expression, we simulated the proportion of patients surviving (number at risk at follow-up time points relative to number at risk at diagnosis) at any given time by plotting and calculating the shift in the baseline OS curve for 789 breast cancer patients (59 death events) using parameters from the output of the multivariate regression model.

4.4. OS Analysis Using Kaplan–Meier Comparisons for Genes Identified in the TCGA Dataset

The KMplotter web tool (https://kmplot.com/analysis/index.php?p=service&cancer=breast, accessed on 15 February 2025) was utilized to access ovarian cancer data from the Affymetrix dataset (N = 1879 patients) for the selection of genes identified from the TCGA dataset (by cross-referencing the KMplotter genes (HR > 1 and p < 0.01)). Patients were grouped based on the median cut-off for gene expression levels to compare the impact of high versus low expression patient sub-groups (Follow-up threshold = 120 months). The web tool reported the hazard ratio (HR) and log-rank p-values [131,132].
Gene expression values were correlated with OS outcomes investigating the impact of mRNA expression of TGFB2 (median cut-off for high and low levels of expression), further stratified into four groups based on gene expression levels of Gene2 mRNA expression levels (median cut-off for high and low levels) in these patients. We defined TGFB2 dependency by visualizing the significant interaction effect from the multivariate Cox proportional hazards models using Kaplan–Meier analysis with 50th percentile cut-offs for TGFB2 and Gene2. This analysis was implemented to show that when comparing high versus low levels of Gene2 expression, the impact on OS was different when comparing high and low levels of TGFB2 mRNA expression or when comparing high versus low levels of TGFB2 mRNA; the OS curves were differentially impacted at low and high levels of Gene2. We identified TGFB2/Gene2 combinations that exhibited significant curve separation between the two arms of the four survival curves (p < 0.05), whereby high levels of TGFB2 and maker gene mRNA expression resulted in the most improved OS outcomes.
We employed two screens to identify TGFB2-dependent markers: 1. in our first screen, the expression of genes that exhibited significant prognostic impact for TGFB2 mRNA expression (p < 0.05) AND Gene2 expression (p < 0.05) AND TGFB2 by Gene2 interaction terms (p < 0.05); and 2. the second screen considered Gene2, which showed significant improvement in OS outcomes when comparing TGFB2high/Gene2high versus TGFB2high/Gene2high-expressing cohorts of patients using Kaplan–Meier analysis.
We identified negative prognostic markers for the whole cohort of breast cancer patients using a Kaplan–Meier analysis of the marker Gene2 that exhibited worse OS times at high levels of Gene2 expression in both TCGA and KMplotter datasets.

4.5. Identifying Prognostically Relevant Signaling Proteins Networked for TGFB2-Dependent Gene2, PAM50, and OncotypeDx Signature Genes Using the STRING Interaction Algorithm

Using the STRING interaction algorithm for putative protein–protein interaction (PPI), networks derived from mRNA levels that exhibited significant prognostic impact on the OS were constructed using the STRING version 12 algorithm (https://string-db.org/cgi/input?sessionId=bR8QWNAdLhxe&input_page_show_search=on, accessed 29 October 2025) to identify candidate hub proteins connecting and/or Gene2 expression that potentially networked interactions between proteins coded by the mRNA [133]. In these diagrams, nodes represent protein identifiers while edges illustrate associations between proteins. Network visualization was achieved by seeding inputs from a list of genes that exhibited significant upregulation in tumor tissue and demonstrated prognostic relevance for TGFB2-dependent Gene2: the gene list from PAM50 [67] and OncotypeDx [71] gene signatures. The edges depicted the confidence level for each association, determined through multiple sources of experimental evidence, including text mining, laboratory experiments, databases, co-expression, neighborhood analysis, gene fusion, and co-occurrence. Interaction edge scores exceeding 0.4 were used to define protein associations, with the thickness of connecting lines denoting score thresholds of 0.4, 0.7, and 0.9. We investigated the network association of the TGFB2 protein with the TGFB2-dependent gene signatures, specifically the PAM50 and OncotypeDx signatures. We identified linking proteins using the capability to identify second-shell proteins with “no more than 20 interactors,” as provided in the “settings” tab of the STRING web portal.

5. Conclusions

For TGFB2-dependent biomarkers, elevated TGFB2 mRNA expression is a prognostic biomarker associated with breast cancer patients that is correlated with improved OS outcomes at high expression levels of GDAP1, TBL1XR1, RNFT1, HACL1, SLC27A2, NLE1, and TXNDC16. This suggests that patients with high levels of TGFB2 and Gene2 were correlated with improved OS and used as markers for future prospective clinical trials. A multivariate analysis revealed that when controlling TGFB2 expression, we identified six genes (ENO1, GLRX2, PLOD1, PRDX4, TAGLN2, and TMED9) that exhibited a negative correlation between mRNA expression and OS across all expression levels in the breast cancer patient, validated with an independent dataset from the KMplotter database. Five of these negative prognostic markers are druggable (ENO1, PLOD1, PRDX4, TAGLN2, and TMED9). The increased HR for patients with high levels of expression of these Gene2s suggests that these five genes could be presented for targeted therapies following confirmation of protein levels and drug action in breast cancer patients. Examination of the protein–protein interaction networks suggested that the correlation of TGFB2-related markers could potentially complement the PAM50 signature in the assessment of OS prognosis in breast cancer patients following validation of the TGFB2/EGFR/MYC protein associations in tumors.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms262311580/s1.

Author Contributions

Evaluations reported in this paper, data compilation and analysis, review and editing: V.T.; conceptualization: V.T., S.Q. and S.R.; methodology, statistical and bioinformatic analyses of the manuscript, data accuracy validation and analysis, original draft preparation: S.Q.; writing—final draft: S.Q., V.T. and S.R.; development of AI software and algorithm tools for PubMed article identification, writing of methods: M.P. and S.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

We utilized log2-transformed transcripts per million (TPM), summarized RNA-seq datafiles (https://toil-xena-hub.s3.us-east-1.amazonaws.com/download/TcgaTargetGtex_rsem_gene_tpm.gz (accessed on 10 November 2025); full metadata) downloaded from the UCSC Xena web platform (https://xenabrowser.net/datapages/, accessed on 25 July 2023). We analyzed clinical metadata (https://www.cbioportal.org/study/summary?id=brca_tcga_pan_can_atlas_2018; accessed on 1 February 2024) and RNA sequencing data (“data_mrna_seq_v2_rsem.txt”: batch normalized from Illumina HiSeq_RNASeqV2 using the RSEM algorithm (transcripts per million (TPM)) from 789 patients diagnosed with breast cancer to compare OS correlates with mRNA expression values.

Conflicts of Interest

Work was funded by Oncotelic Therapeutics as follows: S.Q. as consultant, S.R. as consultant, V.T. as chairman/CEO of Oncotelic Therapeutics, M.P. as employee of Oncotelic Therapeutics, and S.M. as consultant of Oncotelic Therapeutics. Oncotelic Therapeutics had no role in the design of the study, the collection, analysis, or interpretation of the data, the writing of the manuscript, or the decision to publish the results. S.Q. is an independent consultant for Oncotelic Therapeutics and has no shares or any other financial interests in the company. The consultancy requires a rigorous appraisal of the potential use of anti-TGFB2 therapies. Indeed, the bioinformatic analysis carried out in this present study does not warrant the use of anti-TGFB2 therapies in breast cancers. Expression of TGFB2 mRNA could be used as a positive prognostic biomarker in combination of the identified Gene2 to include patients susceptible to standard therapies. This finding provides balance to other published papers by Oncotelic, which suggest the use of anti-TGFB2 therapies for patients with high levels of TGFB2 mRNA in cancers such as gliomas and pancreatic cancers. The TGFB2 drug candidate is still going through clinical development, so there are no commercial conflicts of interest. The bioinformatic efforts are independent and hypothesis-generating and point out potential research opportunities, just like efforts at academic laboratories.

References

  1. Allott, E.H.; Shan, Y.; Chen, M.; Sun, X.; Garcia-Recio, S.; Kirk, E.L.; Olshan, A.F.; Geradts, J.; Earp, H.S.; Carey, L.A.; et al. Bimodal Age Distribution at Diagnosis in Breast Cancer Persists Across Molecular and Genomic Classifications. Breast Cancer Res. Treat. 2020, 179, 185–195. [Google Scholar] [CrossRef]
  2. Guo, L.; Chen, G.; Zhang, W.; Zhou, L.; Xiao, T.; Di, X.; Wang, Y.; Feng, L.; Zhang, K. A High-Risk Luminal A Dominant Breast Cancer Subtype with Increased Mobility. Breast Cancer Res. Treat. 2019, 175, 459–472. [Google Scholar] [CrossRef]
  3. Horr, C.; Buechler, S.A. Breast Cancer Consensus Subtypes: A System for Subtyping Breast Cancer Tumors Based on Gene Expression. NPJ Breast Cancer 2021, 7, 136. [Google Scholar] [CrossRef] [PubMed]
  4. Xu, J.; Qin, S.; Yi, Y.; Gao, H.; Liu, X.; Ma, F.; Guan, M. Delving into the Heterogeneity of Different Breast Cancer Subtypes and the Prognostic Models Utilizing ScRNA-Seq and Bulk RNA-Seq. Int. J. Mol. Sci. 2022, 23, 9936. [Google Scholar] [CrossRef]
  5. Ali, S.; Rehman, M.U.; Yatoo, A.M.; Arafah, A.; Khan, A.; Rashid, S.; Majid, S.; Ali, A.; Ali, N.M. TGF-β Signaling Pathway: Therapeutic Targeting and Potential for Anti-Cancer Immunity. Eur. J. Pharmacol. 2023, 947, 175678. [Google Scholar] [CrossRef]
  6. Cecerska-Heryć, E.; Jerzyk, A.; Goszka, M.; Polikowska, A.; Rachwalska, J.; Serwin, N.; Wojciuk, B.; Dołęgowska, B. TGF-β Signaling in Cancer: Mechanisms of Progression and Therapeutic Targets. Int. J. Mol. Sci. 2025, 26, 7326. [Google Scholar] [CrossRef] [PubMed]
  7. Zhao, M.; Mishra, L.; Deng, C.-X. The Role of TGF-β/SMAD4 Signaling in Cancer. Int. J. Biol. Sci. 2018, 14, 111–123. [Google Scholar] [CrossRef]
  8. Tzavlaki, K.; Moustakas, A. TGF-β Signaling. Biomolecules 2020, 10, 487. [Google Scholar] [CrossRef]
  9. Huang, C.-Y.; Chung, C.-L.; Hu, T.-H.; Chen, J.-J.; Liu, P.-F.; Chen, C.-L. Recent Progress in TGF-β Inhibitors for Cancer Therapy. Biomed. Pharmacother. 2021, 134, 111046. [Google Scholar] [CrossRef] [PubMed]
  10. Qazi, S.; Talebi, Z.; Trieu, V. Transforming Growth Factor Beta 2 (TGFB2) and Interferon Gamma Receptor 2 (IFNGR2) MRNA Levels in the Brainstem Tumor Microenvironment (TME) Significantly Impact Overall Survival in Pediatric DMG Patients. Biomedicines 2024, 12, 191. [Google Scholar] [CrossRef]
  11. Qazi, S.; Trieu, V. TGFB2 MRNA Levels Prognostically Interact with Interferon-Alpha Receptor Activation of IRF9 and IFI27, and an Immune Checkpoint LGALS9 to Impact Overall Survival in Pancreatic Ductal Adenocarcinoma. Int. J. Mol. Sci. 2024, 25, 11221. [Google Scholar] [CrossRef]
  12. Trieu, V.; Maida, A.E.; Qazi, S. Transforming Growth Factor Beta 2 (TGFB2) MRNA Levels, in Conjunction with Interferon-Gamma Receptor Activation of Interferon Regulatory Factor 5 (IRF5) and Expression of CD276/B7-H3, Are Therapeutically Targetable Negative Prognostic Markers in Low-Grade Gliomas. Cancers 2024, 16, 1202. [Google Scholar] [CrossRef]
  13. Figueroa, J.D.; Flanders, K.C.; Garcia-Closas, M.; Anderson, W.F.; Yang, X.R.; Matsuno, R.K.; Duggan, M.A.; Pfeiffer, R.M.; Ooshima, A.; Cornelison, R.; et al. Expression of TGF-β Signaling Factors in Invasive Breast Cancers: Relationships with Age at Diagnosis and Tumor Characteristics. Breast Cancer Res. Treat. 2010, 121, 727–735. [Google Scholar] [CrossRef]
  14. Soufla, G.; Porichis, F.; Sourvinos, G.; Vassilaros, S.; Spandidos, D.A. Transcriptional Deregulation of VEGF, FGF2, TGF-Β1, 2, 3 and Cognate Receptors in Breast Tumorigenesis. Cancer Lett. 2006, 235, 100–113. [Google Scholar] [CrossRef]
  15. Chakravarthy, D.; Green, A.R.; Green, V.L.; Kerin, M.J.; Speirs, V. Expression and Secretion of TGF-Beta Isoforms and Expression of TGF-Beta-Receptors I, II and III in Normal and Neoplastic Human Breast. Int. J. Oncol. 1999, 15, 187–281. [Google Scholar] [CrossRef] [PubMed]
  16. Koumoundourou, D.; Kassimatis, T.; Zolota, V.; Tzorakoeleftherakis, E.; Ravazoula, P.; Vassiliou, V.; Kardamakis, D.; Varakis, J. Prognostic Significance of TGFbeta-1 and PSmad2/3 in Breast Cancer Patients with T1-2,N0 Tumours. Anticancer Res. 2007, 27, 2613–2620. [Google Scholar] [PubMed]
  17. Ciftci, R.; Tas, F.; Yasasever, C.T.; Aksit, E.; Karabulut, S.; Sen, F.; Keskin, S.; Kilic, L.; Yildiz, İ.; Bozbey, H.U.; et al. High Serum Transforming Growth Factor Beta 1 (TGFB1) Level Predicts Better Survival in Breast Cancer. Tumor Biol. 2014, 35, 6941–6948. [Google Scholar] [CrossRef] [PubMed]
  18. Ghellal, A.; Li, C.; Hayes, M.; Byrne, G.; Bundred, N.; Kumar, S. Prognostic Significance of TGF Beta 1 and TGF Beta 3 in Human Breast Carcinoma. Anticancer Res. 2000, 20, 4413–4418. [Google Scholar]
  19. Auvinen, P.; Lipponen, P.; Johansson, R.; Syrjänen, K. Prognostic Significance of TGF-Beta 1 and TGF-Beta 2 Expressions in Female Breast Cancer. Anticancer Res. 1995, 15, 2627–2631. [Google Scholar]
  20. Huang, H.; Li, T.; Ye, G.; Zhao, L.; Zhang, Z.; Mo, D.; Wang, Y.; Zhang, C.; Deng, H.; Li, G.; et al. High Expression of COL10A1 Is Associated with Poor Prognosis in Colorectal Cancer. OncoTargets Ther. 2018, 11, 1571–1581. [Google Scholar] [CrossRef]
  21. Ke, T.-W.; Chang, S.-C.; Yeh, C.-M.; Lin, S.-H.; Yeh, K.-T. Comprehensive Bioinformatic Analysis Reveals Prognostic Significance and Functional Insights of Candidate Gene Expression in Colorectal Cancer. Sci. Rep. 2025, 15, 5659. [Google Scholar] [CrossRef]
  22. Liang, Y.; Xia, W.; Zhang, T.; Chen, B.; Wang, H.; Song, X.; Zhang, Z.; Xu, L.; Dong, G.; Jiang, F. Upregulated Collagen COL10A1 Remodels the Extracellular Matrix and Promotes Malignant Progression in Lung Adenocarcinoma. Front. Oncol. 2020, 10, 573534. [Google Scholar] [CrossRef]
  23. Zhang, M.; Jin, M.; Gao, Z.; Yu, W.; Zhang, W. High COL10A1 Expression Potentially Contributes to Poor Outcomes in Gastric Cancer with the Help of LEF1 and Wnt2. J. Clin. Lab. Anal. 2022, 36, e24612. [Google Scholar] [CrossRef]
  24. Xu, Q.; Zheng, J.; Su, Z.; Chen, B.; Gu, S. COL10A1 Promotes Tumorigenesis by Modulating CD276 in Pancreatic Adenocarcinoma. BMC Gastroenterol. 2023, 23, 397. [Google Scholar] [CrossRef]
  25. Bret, C.; Moreaux, J.; Schved, J.-F.; Hose, D.; Klein, B. SULFs in Human Neoplasia: Implication as Progression and Prognosis Factors. J. Transl. Med. 2011, 9, 72. [Google Scholar] [CrossRef]
  26. Zhang, B.; Luo, D.; Xiang, L.; Chen, J.; Fang, T. Investigating the Anti-Cancer Potential of Sulfatase 1 and Its Underlying Mechanism in Non-Small Cell Lung Cancer. Cytojournal 2024, 21, 52. [Google Scholar] [CrossRef] [PubMed]
  27. Dhanasekaran, R.; Nakamura, I.; Hu, C.; Chen, G.; Oseini, A.M.; Seven, E.S.; Miamen, A.G.; Moser, C.D.; Zhou, W.; van Kuppevelt, T.H.; et al. Activation of the Transforming Growth Factor-β/SMAD Transcriptional Pathway Underlies a Novel Tumor-promoting Role of Sulfatase 1 in Hepatocellular Carcinoma. Hepatology 2015, 61, 1269–1283. [Google Scholar] [CrossRef]
  28. Fang, X.; Chen, D.; Yang, X.; Cao, X.; Cheng, Q.; Liu, K.; Xu, P.; Wang, Y.; Xu, J.; Zhao, S.; et al. Cancer Associated Fibroblasts-Derived SULF1 Promotes Gastric Cancer Metastasis and CDDP Resistance through the TGFBR3-Mediated TGF-β Signaling Pathway. Cell Death Discov. 2024, 10, 111. [Google Scholar] [CrossRef]
  29. Yang, Y.; Ahn, J.; Edwards, N.J.; Benicky, J.; Rozeboom, A.M.; Davidson, B.; Karamboulas, C.; Nixon, K.C.J.; Ailles, L.; Goldman, R. Extracellular Heparan 6-O-Endosulfatases SULF1 and SULF2 in Head and Neck Squamous Cell Carcinoma and Other Malignancies. Cancers 2022, 14, 5553. [Google Scholar] [CrossRef] [PubMed]
  30. Lyu, Y.; Cheng, Y.; Wang, B.; Chen, L.; Zhao, S. Sulfatase 1 Expression in Pancreatic Cancer and Its Correlation with Clinicopathological Features and Postoperative Prognosis. Cancer Biomark. 2018, 22, 701–707. [Google Scholar] [CrossRef] [PubMed]
  31. Hammond, E.; Khurana, A.; Shridhar, V.; Dredge, K. The Role of Heparanase and Sulfatases in the Modification of Heparan Sulfate Proteoglycans within the Tumor Microenvironment and Opportunities for Novel Cancer Therapeutics. Front. Oncol. 2014, 4, 195. [Google Scholar] [CrossRef] [PubMed]
  32. Khurana, A.; Liu, P.; Mellone, P.; Lorenzon, L.; Vincenzi, B.; Datta, K.; Yang, B.; Linhardt, R.J.; Lingle, W.; Chien, J.; et al. HSulf-1 Modulates FGF2- and Hypoxia-Mediated Migration and Invasion of Breast Cancer Cells. Cancer Res. 2011, 71, 2152–2161. [Google Scholar] [CrossRef] [PubMed]
  33. Staub, J.; Chien, J.; Pan, Y.; Qian, X.; Narita, K.; Aletti, G.; Scheerer, M.; Roberts, L.R.; Molina, J.; Shridhar, V. Epigenetic Silencing of HSulf-1 in Ovarian Cancer:Implications in Chemoresistance. Oncogene 2007, 26, 4969–4978. [Google Scholar] [CrossRef]
  34. He, X.; Khurana, A.; Roy, D.; Kaufmann, S.; Shridhar, V. Loss of HSulf-1 Expression Enhances Tumorigenicity by Inhibiting Bim Expression in Ovarian Cancer. Int. J. Cancer 2014, 135, 1783–1789. [Google Scholar] [CrossRef]
  35. Yang, X.; Wei, M.; Huang, Y.; Yang, X.; Yuan, Z.; Huang, J.; Wei, J.; Tian, L. ITGA11, a Prognostic Factor Associated with Immunity in Gastric Adenocarcinoma. Int. J. Gen. Med. 2024, 17, 471–483. [Google Scholar] [CrossRef]
  36. Ando, T.; Kage, H.; Matsumoto, Y.; Zokumasu, K.; Yotsumoto, T.; Maemura, K.; Amano, Y.; Watanabe, K.; Nakajima, J.; Nagase, T.; et al. Integrin A11 in Non–Small Cell Lung Cancer Is Associated with Tumor Progression and Postoperative Recurrence. Cancer Sci. 2020, 111, 200–208. [Google Scholar] [CrossRef]
  37. Primac, I.; Maquoi, E.; Blacher, S.; Heljasvaara, R.; Van Deun, J.; Smeland, H.Y.H.; Canale, A.; Louis, T.; Stuhr, L.; Sounni, N.E.; et al. Stromal Integrin A11 Regulates PDGFRβ Signaling and Promotes Breast Cancer Progression. J. Clin. Investig. 2019, 129, 4609–4628. [Google Scholar] [CrossRef]
  38. Talia, M.; Cesario, E.; Cirillo, F.; Scordamaglia, D.; Di Dio, M.; Zicarelli, A.; Mondino, A.A.; Occhiuzzi, M.A.; De Francesco, E.M.; Belfiore, A.; et al. Cancer-Associated Fibroblasts (CAFs) Gene Signatures Predict Outcomes in Breast and Prostate Tumor Patients. J. Transl. Med. 2024, 22, 597. [Google Scholar] [CrossRef]
  39. Li, M.; Tian, X.; Guo, H.; Xu, X.; Liu, Y.; Hao, X.; Fei, H. A Novel LncRNA-MRNA-MiRNA Signature Predicts Recurrence and Disease-Free Survival in Cervical Cancer. Braz. J. Med. Biol. Res. 2021, 54, e11592. [Google Scholar] [CrossRef] [PubMed]
  40. Decruyenaere, P.; Giuili, E.; Verniers, K.; Anckaert, J.; De Grove, K.; Van der Linden, M.; Deeren, D.; Van Dorpe, J.; Offner, F.; Vandesompele, J. Exploring the Cell-Free Total RNA Transcriptome in Diffuse Large B-Cell Lymphoma and Primary Mediastinal B-Cell Lymphoma Patients as Biomarker Source in Blood Plasma Liquid Biopsies. Front. Oncol. 2023, 13, 1221471. [Google Scholar] [CrossRef]
  41. Takanami, I.; Takeuchi, K.; Watanabe, H.; Yanagawa, T.; Takagishi, K.; Raz, A. Significance Ofautocrine Motility Factor Receptor Gene Expression as a Prognostic Factor in Non-Small-Cell Lung Cancer. Int. J. Cancer 2001, 95, 384–387. [Google Scholar] [CrossRef]
  42. Huang, Z.; Zhang, N.; Zha, L.; Mao, H.-C.; Chen, X.; Xiang, J.-F.; Zhang, H.; Wang, Z.-W. Aberrant Expression of the Autocrine Motility Factor Receptor Correlates with Poor Prognosis and Promotes Metastasis in Gastric Carcinoma. Asian Pac. J. Cancer Prev. 2014, 15, 989–997. [Google Scholar] [CrossRef]
  43. Jiang, W.G.; Raz, A.; Douglas-Jones, A.; Mansel, R.E. Expression of Autocrine Motility Factor (AMF) and Its Receptor, AMFR, in Human Breast Cancer. J. Histochem. Cytochem. 2006, 54, 231–241. [Google Scholar] [CrossRef]
  44. Halwani, Y.; Kojic, L.D.; Chan, S.K.; Phang, T.P.; Masoudi, H.; Jones, S.J.M.; Nabi, I.R.; Wiseman, S.M. Prognostic Significance of Autocrine Motility Factor Receptor Expression by Colorectal Cancer and Lymph Node Metastases. Am. J. Surg. 2015, 209, 884–889. [Google Scholar] [CrossRef]
  45. Marco, A. Evolutionary and Structural Analyses of GDAP1, Involved in Charcot-Marie-Tooth Disease, Characterize a Novel Class of Glutathione Transferase-Related Genes. Mol. Biol. Evol. 2003, 21, 176–187. [Google Scholar] [CrossRef] [PubMed]
  46. Magouliotis, D.E.; Sakellaridis, N.; Dimas, K.; Tasiopoulou, V.S.; Svokos, K.A.; Svokos, A.A.; Zacharoulis, D. In Silico Transcriptomic Analysis of the Chloride Intracellular Channels (CLIC) Interactome Identifies a Molecular Panel of Seven Prognostic Markers in Patients with Pancreatic Ductal Adenocarcinoma. Curr. Genom. 2020, 21, 119–127. [Google Scholar] [CrossRef] [PubMed]
  47. Yu, X.; Chen, C.; Hu, Y.; Li, K.; Zhang, Y.; Chen, Z.; Nie, D.; Gao, R.; Huang, Y.; Zhong, M.; et al. High Expression of LOC541471, GDAP1, SOD1, and STK25 Is Associated with Poor Overall Survival of Patients with Acute Myeloid Leukemia. Cancer Med. 2023, 12, 9055–9067. [Google Scholar] [CrossRef]
  48. Shen, J.; Shu, M.; Xie, S.; Yan, J.; Pan, K.; Chen, S.; Li, X. A Six-Gene Prognostic Risk Prediction Model In Hepatitis B Virus-Associated Hepatocellular Carcinoma. Clin. Investig. Med. 2021, 44, E32–E44. [Google Scholar] [CrossRef]
  49. Wang, Y.; Chen, X.; Li, Y.; Zhang, Z.; Xia, L.; Jiang, J.; Chai, Y.; Wang, Z.; Wan, Y.; Li, T.; et al. SLC27A2 Is a Potential Immune Biomarker for Hematological Tumors and Significantly Regulates the Cell Cycle Progression of Diffuse Large B-Cell Lymphoma. BMC Med. Genom. 2024, 17, 105. [Google Scholar] [CrossRef] [PubMed]
  50. Zhang, B.; Zhang, Y.; Chang, K.; Hou, N.; Fan, P.; Ji, C.; Liu, L.; Wang, Z.; Li, R.; Wang, Y.; et al. Risk Assessment Model Based on Nucleotide Metabolism-Related Genes Highlights SLC27A2 as a Potential Therapeutic Target in Breast Cancer. J. Cancer Res. Clin. Oncol. 2024, 150, 258. [Google Scholar] [CrossRef]
  51. Feng, K.; Ma, R.; Li, H.; Yin, K.; Du, G.; Chen, X.; Liu, Z.; Yin, D. Upregulated SLC27A2/FATP2 in Differentiated Thyroid Carcinoma Promotes Tumor Proliferation and Migration. J. Clin. Lab. Anal. 2022, 36, e24148. [Google Scholar] [CrossRef] [PubMed]
  52. Lu, L.; Li, J.; Zheng, Y.; Luo, L.; Huang, Y.; Hu, J.; Chen, Y. High Expression of SLC27A2 Predicts Unfavorable Prognosis and Promotes Inhibitory Immune Infiltration in Acute Lymphoblastic Leukemia. Transl. Oncol. 2024, 45, 101952. [Google Scholar] [CrossRef]
  53. Xu, N.; Xiao, W.; Meng, X.; Li, W.; Wang, X.; Zhang, X.; Yang, H. Up-Regulation of SLC27A2 Suppresses the Proliferation and Invasion of Renal Cancer by Down-Regulating CDK3-Mediated EMT. Cell Death Discov. 2022, 8, 351. [Google Scholar] [CrossRef]
  54. Kocatürk, B. Identification of Thioredoxin Domain Containing Family Members’ Expression Pattern and Prognostic Value in Diffuse Gliomas via in Silico Analysis. Cancer Med. 2023, 12, 3830–3844. [Google Scholar] [CrossRef]
  55. Li, X.; Liang, W.; Liu, J.; Lin, C.; Wu, S.; Song, L.; Yuan, Z. Transducin (β)-like 1 X-Linked Receptor 1 Promotes Proliferation and Tumorigenicity in Human Breast Cancer via Activation of Beta-Catenin Signaling. Breast Cancer Res. 2014, 16, 465. [Google Scholar] [CrossRef] [PubMed]
  56. Du, R.; Li, K.; Guo, K.; Chen, Z.; Zhao, X.; Han, L.; Bian, H. Two Decades of a Protooncogene TBL1XR1: From a Transcription Modulator to Cancer Therapeutic Target. Front. Oncol. 2024, 14, 1309687. [Google Scholar] [CrossRef] [PubMed]
  57. Ren, Z.; Ni, F.; Zhang, T.; Yuan, X.; Li, J.; Xiao, W. Knockdown of NLE1 Inhibits Development of Malignant Melanoma In Vitro and In Vivo NLE1 Promotes Development of Malignant Melanoma. Exp. Cell Res. 2021, 404, 112636, Erratum in Exp. Cell Res. 2021, 405, 112733. [Google Scholar] [CrossRef]
  58. Wang, Y.; Xu, Z. The Effect of Up-Regulation of NLE1 Gene Expression on the Invasion and Migration of Colon Cancer Cells and Its Mechanism. J. King Saud Univ. Sci. 2023, 35, 102669. [Google Scholar] [CrossRef]
  59. Xu, P.; Wang, L.; Mo, B.; Xie, X.; Hu, R.; Jiang, L.; Hu, F.; Ding, F.; Xiao, H. Identification of NLE1/CDK1 Axis as Key Regulator in the Development and Progression of Non-Small Cell Lung Cancer. Front. Oncol. 2023, 12, 985827. [Google Scholar] [CrossRef]
  60. Zhang, R.; Liu, D.; Feng, X.; Yang, Z.; Zhang, K. WD40 Protein NLE1 as a Novel Diagnostic Biomarker Promoting Hepatocellular Carcinoma Proliferation. Clin. Med. Insights Oncol. 2025, 19, 11795549251348902. [Google Scholar] [CrossRef]
  61. Ungefroren, H.; Lenschow, W.; Chen, W.-B.; Faendrich, F.; Kalthoff, H. Regulation of Biglycan Gene Expression by Transforming Growth Factor-β Requires MKK6-P38 Mitogen-Activated Protein Kinase Signaling Downstream of Smad Signaling. J. Biol. Chem. 2003, 278, 11041–11049. [Google Scholar] [CrossRef] [PubMed]
  62. Ungefroren, H.; Groth, S.; Ruhnke, M.; Kalthoff, H.; Fändrich, F. Transforming Growth Factor-β (TGF-β) Type I Receptor/ALK5-Dependent Activation of the GADD45β Gene Mediates the Induction of Biglycan Expression by TGF-β. J. Biol. Chem. 2005, 280, 2644–2652. [Google Scholar] [CrossRef]
  63. Verrecchia, F.; Chu, M.-L.; Mauviel, A. Identification of Novel TGF-β/Smad Gene Targets in Dermal Fibroblasts Using a Combined CDNA Microarray/Promoter Transactivation Approach. J. Biol. Chem. 2001, 276, 17058–17062. [Google Scholar] [CrossRef]
  64. Zhao, L.; Wang, P.; Sun, L.; Ma, W.; Yu, L. SP1/COL1A2/ZEB1 Axis Promotes TGF-Β2-Induced Lens Epithelial Cell Proliferation, Migration, Invasion and EMT Process. Exp. Eye Res. 2025, 251, 110220. [Google Scholar] [CrossRef]
  65. Han, Y.-L.; Luo, D.; Habaxi, K.; Tayierjiang, J.; Zhao, W.; Wang, W.; Aikebaier, W.; Wang, L. COL5A2 Inhibits the TGF-β and Wnt/β-Catenin Signaling Pathways to Inhibit the Invasion and Metastasis of Osteosarcoma. Front. Oncol. 2022, 12, 813809. [Google Scholar] [CrossRef]
  66. Nan, P.; Dong, X.; Bai, X.; Lu, H.; Liu, F.; Sun, Y.; Zhao, X. Tumor-Stroma TGF-Β1-THBS2 Feedback Circuit Drives Pancreatic Ductal Adenocarcinoma Progression via Integrin Avβ3/CD36-Mediated Activation of the MAPK Pathway. Cancer Lett. 2022, 528, 59–75. [Google Scholar] [CrossRef]
  67. Bastien, R.R.; Rodríguez-Lescure, Á.; Ebbert, M.T.; Prat, A.; Munárriz, B.; Rowe, L.; Miller, P.; Ruiz-Borrego, M.; Anderson, D.; Lyons, B.; et al. PAM50 Breast Cancer Subtyping by RT-QPCR and Concordance with Standard Clinical Molecular Markers. BMC Med. Genom. 2012, 5, 44. [Google Scholar] [CrossRef]
  68. Ohnstad, H.O.; Borgen, E.; Falk, R.S.; Lien, T.G.; Aaserud, M.; Sveli, M.A.T.; Kyte, J.A.; Kristensen, V.N.; Geitvik, G.A.; Schlichting, E.; et al. Prognostic Value of PAM50 and Risk of Recurrence Score in Patients with Early-Stage Breast Cancer with Long-Term Follow-Up. Breast Cancer Res. 2017, 19, 120. [Google Scholar] [CrossRef] [PubMed]
  69. Cichon, M.A.; Moruzzi, M.E.; Shqau, T.A.; Miller, E.; Mehner, C.; Ethier, S.P.; Copland, J.A.; Radisky, E.S.; Radisky, D.C. MYC Is a Crucial Mediator of TGFβ-Induced Invasion in Basal Breast Cancer. Cancer Res. 2016, 76, 3520–3530. [Google Scholar] [CrossRef] [PubMed]
  70. Zhao, Y.; Ma, J.; Fan, Y.; Wang, Z.; Tian, R.; Ji, W.; Zhang, F.; Niu, R. TGF-β Transactivates EGFR and Facilitates Breast Cancer migration and Invasion Through Canonical Smad3 and ERK/Sp1 Signaling Pathways. Mol. Oncol. 2018, 12, 305–321. [Google Scholar] [CrossRef]
  71. Kalinsky, K.; Barlow, W.E.; Gralow, J.R.; Meric-Bernstam, F.; Albain, K.S.; Hayes, D.F.; Lin, N.U.; Perez, E.A.; Goldstein, L.J.; Chia, S.K.L.; et al. 21-Gene Assay to Inform Chemotherapy Benefit in Node-Positive Breast Cancer. N. Engl. J. Med. 2021, 385, 2336–2347. [Google Scholar] [CrossRef]
  72. Sun, X.; Bernhardt, S.M.; Glynn, D.J.; Hodson, L.J.; Woolford, L.; Evdokiou, A.; Yan, C.; Du, H.; Robertson, S.A.; Ingman, W.V. Attenuated TGFB Signalling in Macrophages Decreases Susceptibility to DMBA-Induced Mammary Cancer in Mice. Breast Cancer Res. 2021, 23, 39. [Google Scholar] [CrossRef]
  73. Chow, A.; Arteaga, C.L.; Wang, S.E. When Tumor Suppressor TGFβ Meets the HER2 (ERBB2) Oncogene. J. Mammary Gland. Biol. Neoplasia 2011, 16, 81–88. [Google Scholar] [CrossRef]
  74. Shi, Y.-Y.; Chen, X.-L.; Chen, Q.-X.; Yang, Y.-Z.; Zhou, M.; Ren, Y.-X.; Tang, L.-Y.; Ren, Z.-F. Association of Enolase-1 with Prognosis and Immune Infiltration in Breast Cancer by Clinical Stage. J. Inflamm. Res. 2023, 16, 493–503. [Google Scholar] [CrossRef] [PubMed]
  75. Liu, Y.; Liao, L.; An, C.; Wang, X.; Li, Z.; Xu, Z.; Liu, J.; Liu, S. α-Enolase Lies Downstream of MTOR/HIF1α and Promotes Thyroid Carcinoma Progression by Regulating CST1. Front. Cell Dev. Biol. 2021, 9, 670019. [Google Scholar] [CrossRef] [PubMed]
  76. Xu, W.; Yang, W.; Wu, C.; Ma, X.; Li, H.; Zheng, J. Enolase 1 Correlated with Cancer Progression and Immune-Infiltrating in Multiple Cancer Types: A Pan-Cancer Analysis. Front. Oncol. 2021, 10, 593706. [Google Scholar] [CrossRef] [PubMed]
  77. Huang, C.K.; Sun, Y.; Lv, L.; Ping, Y. ENO1 and Cancer. Mol. Ther. Oncolytics 2022, 24, 288–298. [Google Scholar] [CrossRef]
  78. Almaguel, F.A.; Sanchez, T.W.; Ortiz-Hernandez, G.L.; Casiano, C.A. Alpha-Enolase: Emerging Tumor-Associated Antigen, Cancer Biomarker, and Oncotherapeutic Target. Front. Genet. 2021, 11, 614726. [Google Scholar] [CrossRef]
  79. Novelli, F. Alpha-Enolase i ENO1 i a Potential Target in Novel Immunotherapies. Front. Biosci. 2017, 22, 4526. [Google Scholar] [CrossRef]
  80. Chen, M.-L.; Yuan, T.-T.; Chuang, C.-F.; Huang, Y.-T.; Chung, I.-C.; Huang, W.-C. A Novel Enolase-1 Antibody Targets Multiple Interacting Players in the Tumor Microenvironment of Advanced Prostate Cancer. Mol. Cancer Ther. 2022, 21, 1337–1347. [Google Scholar] [CrossRef]
  81. Li, Y.; Liu, L.; Li, B. Role of ENO1 and Its Targeted Therapy in Tumors. J. Transl. Med. 2024, 22, 1025. [Google Scholar] [CrossRef]
  82. Lin, Y.; Zhang, W.; Liu, L.; Li, W.; Li, Y.; Li, B. ENO1 Promotes OSCC Migration and Invasion by Orchestrating IL-6 Secretion from Macrophages via a Positive Feedback Loop. Int. J. Mol. Sci. 2023, 24, 737. [Google Scholar] [CrossRef] [PubMed]
  83. Xu, X.; Chen, B.; Zhu, S.; Zhang, J.; He, X.; Cao, G.; Chen, B. Hyperglycemia Promotes Snail-Induced Epithelial–Mesenchymal Transition of Gastric Cancer via Activating ENO1 Expression. Cancer Cell Int. 2019, 19, 344. [Google Scholar] [CrossRef]
  84. Zhao, M.; Fang, W.; Wang, Y.; Guo, S.; Shu, L.; Wang, L.; Chen, Y.; Fu, Q.; Liu, Y.; Hua, S.; et al. Enolase-1 Is a Therapeutic Target in Endometrial Carcinoma. Oncotarget 2015, 6, 15610–15627. [Google Scholar] [CrossRef]
  85. Ji, H.; Wang, J.; Guo, J.; Li, Y.; Lian, S.; Guo, W.; Yang, H.; Kong, F.; Zhen, L.; Guo, L.; et al. Progress in the Biological Function of Alpha-Enolase. Anim. Nutr. 2016, 2, 12–17. [Google Scholar] [CrossRef]
  86. Chung, I.-C.; Huang, W.-C.; Huang, Y.-T.; Chen, M.-L.; Tsai, A.-W.; Wu, P.-Y.; Yuan, T.-T. Unrevealed Roles of Extracellular Enolase-1 (ENO1) in Promoting Glycolysis and Pro-Cancer Activities in Multiple Myeloma via Hypoxia-Inducible Factor 1α. Oncol. Rep. 2023, 50, 205. [Google Scholar] [CrossRef]
  87. Song, Q.; Zhang, K.; Sun, T.; Xu, C.; Zhao, W.; Zhang, Z. Knockout of ENO1 Leads to Metabolism Reprogramming and Tumor Retardation in Pancreatic Cancer. Front. Oncol. 2023, 13, 1119886. [Google Scholar] [CrossRef]
  88. Su, Z.; You, L.; He, Y.; Chen, J.; Zhang, G.; Liu, Z. Multi-Omics Reveals the Role of ENO1 in Bladder Cancer and Constructs an Epithelial-Related Prognostic Model to Predict Prognosis and Efficacy. Sci. Rep. 2024, 14, 2189. [Google Scholar] [CrossRef] [PubMed]
  89. Jung, D.-W.; Kim, W.-H.; Park, S.-H.; Lee, J.; Kim, J.; Su, D.; Ha, H.-H.; Chang, Y.-T.; Williams, D.R. A Unique Small Molecule Inhibitor of Enolase Clarifies Its Role in Fundamental Biological Processes. ACS Chem. Biol. 2013, 8, 1271–1282. [Google Scholar] [CrossRef] [PubMed]
  90. Shen, C.; Liu, J.; Hu, D.; Liu, C.; Xie, F.; Wang, Y. Tumor-Intrinsic ENO1 Inhibition Promotes Antitumor Immune Response and Facilitates the Efficacy of Anti-PD-L1 Immunotherapy in Bladder Cancer. J. Exp. Clin. Cancer Res. 2025, 44, 207. [Google Scholar] [CrossRef]
  91. León-Letelier, R.A.; Sevillano-Mantas, A.M.; Chen, Y.; Park, S.; Vykoukal, J.; Fahrmann, J.F.; Ostrin, E.J.; Garrett, C.; Dou, R.; Cai, Y.; et al. Citrullinated ENO1 Vaccine Enhances PD-1 Blockade in Mice Implanted with Murine Triple-Negative Breast Cancer Cells. Vaccines 2025, 13, 629. [Google Scholar] [CrossRef]
  92. Li, J.; Tang, X.; Wen, X.; Ren, X.; Zhang, H.; Du, Y.; Lu, J. Mitochondrial Glrx2 Knockout Augments Acetaminophen-Induced Hepatotoxicity in Mice. Antioxidants 2022, 11, 1643. [Google Scholar] [CrossRef]
  93. Brzozowa-Zasada, M.; Piecuch, A.; Bajdak-Rusinek, K.; Gołąbek, K.; Michalski, M.; Janelt, K.; Matysiak, N. Glutaredoxin 2 Protein (Grx2) as an Independent Prognostic Factor Associated with the Survival of Colon Adenocarcinoma Patients. Int. J. Mol. Sci. 2024, 25, 1060. [Google Scholar] [CrossRef] [PubMed]
  94. Li, J.; Wang, Z.; Wang, T. Machine-Learning Prediction of a Novel Diagnostic Model Using Mitochondria-Related Genes for Patients with Bladder Cancer. Sci. Rep. 2024, 14, 9282. [Google Scholar] [CrossRef] [PubMed]
  95. Chang, Y.; Li, G.; Zhai, Y.; Huang, L.; Feng, Y.; Wang, D.; Zhang, W.; Hu, H. Redox Regulator GLRX Is Associated with Tumor Immunity in Glioma. Front. Immunol. 2020, 11, 580934. [Google Scholar] [CrossRef]
  96. Jiang, H.; Guo, W.; Yuan, S.; Song, L. PLOD1 Is a Prognostic Biomarker and Mediator of Proliferation and Invasion in Osteosarcoma. Biomed. Res. Int. 2020, 2020, 3418398. [Google Scholar] [CrossRef]
  97. Wang, Z.; Shi, Y.; Ying, C.; Jiang, Y.; Hu, J. Hypoxia-Induced PLOD1 Overexpression Contributes to the Malignant Phenotype of Glioblastoma via NF-ΚB Signaling. Oncogene 2021, 40, 1458–1475. [Google Scholar] [CrossRef] [PubMed]
  98. Zhang, Y.; Wu, Y.; Su, X. PLOD1 Promotes Cell Growth and Aerobic Glycolysis by Regulating the SOX9/PI3K/Akt/MTOR Signaling Pathway in Gastric Cancer. Front. Biosci. 2021, 26, 322–334. [Google Scholar] [CrossRef]
  99. Yang, Z.; Zhou, Z.; Yang, M. Lysyl Hydroxylase PLOD1 Enhances Actin Network to Promote Confined Migration of Hepatocellular Carcinoma Cells via Binding with Septin2. J. Clin. Oncol. 2022, 40, e16111. [Google Scholar] [CrossRef]
  100. Wang, D.-D.; Li, L.; Fu, Y.-Q.; Yang, S.-J.; Chen, X.; Hou, J.-C.; Zhang, Q.; Tian, X.-X.; Tang, J.-H.; Zhang, J.; et al. Systematic Characterization of the Expression, Prognosis and Immune Characteristics of PLOD Family Genes in Breast Cancer. Aging 2024, 16, 11434–11445. [Google Scholar] [CrossRef]
  101. Zhang, J.; Tian, Y.; Mo, S.; Fu, X. Overexpressing PLOD Family Genes Predict Poor Prognosis in Pancreatic Cancer. Int. J. Gen. Med. 2022, 15, 3077–3096. [Google Scholar] [CrossRef] [PubMed]
  102. Li, S.-S.; Lian, Y.-F.; Huang, Y.-L.; Huang, Y.-H.; Xiao, J. Overexpressing PLOD Family Genes Predict Poor Prognosis in Gastric Cancer. J. Cancer 2020, 11, 121–131. [Google Scholar] [CrossRef]
  103. Yamada, Y.; Kato, M.; Arai, T.; Sanada, H.; Uchida, A.; Misono, S.; Sakamoto, S.; Komiya, A.; Ichikawa, T.; Seki, N. Aberrantly Expressed PLOD1 Promotes Cancer Aggressiveness in Bladder Cancer: A Potential Prognostic Marker and Therapeutic Target. Mol. Oncol. 2019, 13, 1898–1912. [Google Scholar] [CrossRef]
  104. Zhai, Z.; Wang, S.; Cao, Y.; Liu, J.; Zhao, Q.; Ji, Y.; Yang, X.; Tang, X.; Ma, J.; Du, P. Pan-Cancer Analysis Reveals the Potential of PLOD1 as a Prognostic and Immune Biomarker for Human Cancer. Biomedicines 2024, 12, 2653. [Google Scholar] [CrossRef]
  105. Yeh, T.-L.; Leissing, T.M.; Abboud, M.I.; Thinnes, C.C.; Atasoylu, O.; Holt-Martyn, J.P.; Zhang, D.; Tumber, A.; Lippl, K.; Lohans, C.T.; et al. Molecular and Cellular Mechanisms of HIF Prolyl Hydroxylase Inhibitors in Clinical Trials. Chem. Sci. 2017, 8, 7651–7668. [Google Scholar] [CrossRef]
  106. Jia, W.; Chen, P.; Cheng, Y. PRDX4 and Its Roles in Various Cancers. Technol. Cancer Res. Treat. 2019, 18, 1533033819864313. [Google Scholar] [CrossRef]
  107. Ding, N.; Jiang, H.; Thapa, P.; Hao, Y.; Alshahrani, A.; Allison, D.; Izumi, T.; Rangnekar, V.M.; Liu, X.; Wei, Q. Peroxiredoxin IV Plays a Critical Role in Cancer Cell Growth and Radioresistance Through the Activation of the Akt/GSK3 Signaling Pathways. J. Biol. Chem. 2022, 298, 102123. [Google Scholar] [CrossRef]
  108. Zhang, L.; Wu, K.; Hou, Y.; Li, X. Validation of the Interaction between PRDX4 and TXNDC5 in Gastric Cancer and the Significance of the PRDX4 Gene in Gastric Cancer Based on a Data Mining Analysis. Transl. Cancer Res. 2024, 13, 81–101. [Google Scholar] [CrossRef]
  109. Zhou, H.; Li, L.; Chen, J.; Hou, S.; Zhou, T.; Xiong, Y. Expression and Prognostic Value of PRDX Family in Colon Adenocarcinoma by Integrating Comprehensive Analysis and In Vitro and In Vivo Validation. Front. Oncol. 2023, 13, 1136738. [Google Scholar] [CrossRef] [PubMed]
  110. Zheng, J.; Guo, X.; Nakamura, Y.; Zhou, X.; Yamaguchi, R.; Zhang, J.; Ishigaki, Y.; Uramoto, H.; Yamada, S. Overexpression of PRDX4 Modulates Tumor Microenvironment and Promotes Urethane-Induced Lung Tumorigenesis. Oxid. Med. Cell Longev. 2020, 2020, 8262730. [Google Scholar] [CrossRef] [PubMed]
  111. Zhou, J.-W.; Bai, Y.; Guo, J.-Q.; Li, Y.-Y.; Liu, Y.-F.; Liang, C.; Xing, Y.-R.; Guo, H.-L.; Qi, T.-X.; Wu, J.; et al. Peroxiredoxin 4 as a Switch Regulating PTEN/AKT Axis in Alveolar Macrophages Activation. Signal Transduct. Target. Ther. 2025, 10, 352. [Google Scholar] [CrossRef] [PubMed]
  112. Kim, H.-R.; Park, J.-S.; Karabulut, H.; Yasmin, F.; Jun, C.-D. Transgelin-2: A Double-Edged Sword in Immunity and Cancer Metastasis. Front. Cell Dev. Biol. 2021, 9, 606149. [Google Scholar] [CrossRef]
  113. Jeon, B.-N.; Kim, H.-R.; Chung, Y.S.; Na, B.-R.; Park, H.; Hong, C.; Fatima, Y.; Oh, H.; Kim, C.-H.; Jun, C.-D. Actin Stabilizer TAGLN2 Potentiates Adoptive T Cell Therapy by Boosting the Inside-Out Costimulation via Lymphocyte Function-Associated Antigen-1. Oncoimmunology 2018, 7, e1500674. [Google Scholar] [CrossRef]
  114. Yang, L.; Hong, Q.; Xu, S.; Kuang, X.; Di, G.; Liu, G.; Wu, J.; Shao, Z.; Yu, S. Downregulation of Transgelin 2 Promotes Breast Cancer Metastasis by Activating the Reactive Oxygen Species/Nuclear Factor-κB Signaling Pathway. Mol. Med. Rep. 2019, 20, 4045–4258. [Google Scholar] [CrossRef]
  115. Pan, T.; Wang, S.; Wang, Z. An Integrated Analysis Identified TAGLN2 As an Oncogene Indicator Related to Prognosis and Immunity in Pan-Cancer. J. Cancer 2023, 14, 1809–1836. [Google Scholar] [CrossRef]
  116. Zhuo, H.; Hou, J.; Hong, Z.; Yu, S.; Peng, H.; Zhang, L.; Xie, W.; Hong, X. TAGLN2 Induces Resistance Signature ISGs by Activating AKT-YBX1 Signal with Dual Pathways and Mediates the IFN-Related DNA Damage Resistance in Gastric Cancer. Cell Death Dis. 2024, 15, 608. [Google Scholar] [CrossRef]
  117. Wang, L.; Tan, H.; Huang, Y.; Guo, M.; Dong, Y.; Liu, C.; Zhao, H.; Liu, Z. TAGLN2 Promotes Papillary Thyroid Carcinoma Invasion via the Rap1/PI3K/AKT Axis. Endocr. Relat. Cancer 2023, 30, e210352. [Google Scholar] [CrossRef]
  118. Zhao, Z.; Lu, L.; Li, W. TAGLN2 Promotes the Proliferation, Invasion, Migration and Epithelial-mesenchymal Transition of Colorectal Cancer Cells by Activating STAT3 Signaling through ANXA2. Oncol. Lett. 2021, 22, 737. [Google Scholar] [CrossRef]
  119. Jin, H.; Zheng, W.; Hou, J.; Peng, H.; Zhuo, H. An Essential NRP1-Mediated Role for Tagln2 in Gastric Cancer Angiogenesis. Front. Oncol. 2021, 11, 653246. [Google Scholar] [CrossRef] [PubMed]
  120. Kumar, A.; Rajasekera, P.; Becker, V.; Biehn, S.; Beyer, S.; McElroy, J.; Becker, A.; Johnson, B.; Cui, T.; Sebastian, E.; et al. Small Molecule Inhibitors Targeting Transgelin-2-Actin Interaction for Therapeutic Intervention in Glioblastoma. Int. J. Radiat. Oncol. Biol. Phys. 2022, 114, e72–e73. [Google Scholar] [CrossRef]
  121. Zhou, L.; Li, H.; Yao, H.; Dai, X.; Gao, P.; Cheng, H. TMED Family Genes and Their Roles in Human Diseases. Int. J. Med. Sci. 2023, 20, 1732–1743. [Google Scholar] [CrossRef]
  122. Wang, H.; Wang, Y.; Tan, P.; Liu, Y.; Zhou, S.; Ma, W. Prognostic Value and Anti-Tumor Immunity Role of TMED9 in Pan-Cancer: A Bioinformatics Study. Transl. Cancer Res. 2024, 13, 5429–5445. [Google Scholar] [CrossRef]
  123. Yang, Y.-C.; Chien, M.-H.; Lai, T.-C.; Tung, M.-C.; Jan, Y.-H.; Chang, W.-M.; Jung, S.-M.; Chen, M.-H.; Yeh, C.-N.; Hsiao, M. Proteomics-Based Identification of TMED9 Is Linked to Vascular Invasion and Poor Prognoses in Patients with Hepatocellular Carcinoma. J. Biomed. Sci. 2021, 28, 29. [Google Scholar] [CrossRef]
  124. HAN, G.H.; YUN, H.; CHUNG, J.-Y.; KIM, J.-H.; CHO, H. TMED9 Expression Level as a Biomarker of Epithelial Ovarian Cancer Progression and Prognosis. Cancer Genom. Proteom. 2022, 19, 692–702. [Google Scholar] [CrossRef]
  125. Fang, Z.; Song, Y.-X.; Wo, G.-Q.; Zhou, H.-L.; Li, L.; Yang, S.-Y.; Chen, X.; Zhang, J.; Tang, J.-H. Screening of the Novel Immune-Suppressive Biomarkers of TMED Family and Whether Knockdown of TMED2/3/4/9 Inhibits Cell Migration and Invasion in Breast Cancer. Ann. Transl. Med. 2022, 10, 1280. [Google Scholar] [CrossRef]
  126. Mishra, S.; Bernal, C.; Silvano, M.; Anand, S.; Ruiz i Altaba, A. The Protein Secretion Modulator TMED9 Drives CNIH4/TGFα/GLI Signaling Opposing TMED3-WNT-TCF to Promote Colon Cancer Metastases. Oncogene 2019, 38, 5817–5837. [Google Scholar] [CrossRef] [PubMed]
  127. Ullah, M.A.; Tabassum, T.; Farzana, M.; Moin, A.T.; Zohora, U.S.; Rahman, M.S. Expression Analysis, Molecular Characterization and Prognostic Evaluation on TMED4 and TMED9 Gene Expression in Glioma. Biomed. Signal Process. Control. 2022, 78, 103922. [Google Scholar] [CrossRef]
  128. Dvela-Levitt, M.; Kost-Alimova, M.; Emani, M.; Kohnert, E.; Thompson, R.; Sidhom, E.-H.; Rivadeneira, A.; Sahakian, N.; Roignot, J.; Papagregoriou, G.; et al. Small Molecule Targets TMED9 and Promotes Lysosomal Degradation to Reverse Proteinopathy. Cell 2019, 178, 521–535.e23. [Google Scholar] [CrossRef]
  129. Goldman, M.J.; Craft, B.; Hastie, M.; Repečka, K.; McDade, F.; Kamath, A.; Banerjee, A.; Luo, Y.; Rogers, D.; Brooks, A.N.; et al. Visualizing and Interpreting Cancer Genomics Data via the Xena Platform. Nat. Biotechnol. 2020, 38, 675–678. [Google Scholar] [CrossRef]
  130. Vivian, J.; Rao, A.A.; Nothaft, F.A.; Ketchum, C.; Armstrong, J.; Novak, A.; Pfeil, J.; Narkizian, J.; Deran, A.D.; Musselman-Brown, A.; et al. Toil Enables Reproducible, Open Source, Big Biomedical Data Analyses. Nat. Biotechnol. 2017, 35, 314–316. [Google Scholar] [CrossRef]
  131. Denkert, C.; Budczies, J.; Darb-Esfahani, S.; Györffy, B.; Sehouli, J.; Könsgen, D.; Zeillinger, R.; Weichert, W.; Noske, A.; Buckendahl, A.; et al. A Prognostic Gene Expression Index in Ovarian Cancer—Validation across Different Independent Data Sets. J. Pathol. 2009, 218, 273–280. [Google Scholar] [CrossRef] [PubMed]
  132. Pérez-Pena, J.; Tibor Fekete, J.; Páez, R.; Baliu-Piqué, M.; García-Saenz, J.Á.; García-Barberán, V.; Manzano, A.; Pérez-Segura, P.; Esparis-Ogando, A.; Pandiella, A.; et al. A Transcriptomic Immunologic Signature Predicts Favorable Outcome in Neoadjuvant Chemotherapy Treated Triple Negative Breast Tumors. Front. Immunol. 2019, 10, 2802. [Google Scholar] [CrossRef] [PubMed]
  133. Szklarczyk, D.; Kirsch, R.; Koutrouli, M.; Nastou, K.; Mehryary, F.; Hachilif, R.; Gable, A.L.; Fang, T.; Doncheva, N.T.; Pyysalo, S.; et al. The STRING Database in 2023: Protein–Protein Association Networks and Functional Enrichment Analyses for Any Sequenced Genome of Interest. Nucleic Acids Res. 2023, 51, D638–D646. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Upregulated genes in tumor tissues with significant prognostic impacts of TGFB2 mRNA, Gene2, and significant statistical interaction for TGFB2 and Gene2 expression. Multivariate analyses utilized the Cox proportional hazards model to assess the individual effects of TGFB2 and Gene2 mRNA expression levels. A total of 111 genes had a significant effect of the TGFB2 by Gene2 interaction term and exhibited significant upregulation in tumor tissues (p < 0.0001, FDR < 0.001). (A) The cluster figure shows tumor tissue expression mean centered to normal tissues for 11 out of the 111 genes examined (log2 (tumor/normal) TPM), color-coded from blue to red, depicting a decrease to an increase in expression in tumor tissues, respectively; they were further filtered using significant hazard ratio calculations for TGFB2 mRNA expression (p < 0.05) AND Gene2 expression (p < 0.05) AND TGFB2 by Gene2 interaction terms (p < 0.05). (B) The table depicts a comparison of mRNA expression values, comparing normal (n = 179) versus tumor samples (n = 786 evaluable samples), ordered by fold increase.
Figure 1. Upregulated genes in tumor tissues with significant prognostic impacts of TGFB2 mRNA, Gene2, and significant statistical interaction for TGFB2 and Gene2 expression. Multivariate analyses utilized the Cox proportional hazards model to assess the individual effects of TGFB2 and Gene2 mRNA expression levels. A total of 111 genes had a significant effect of the TGFB2 by Gene2 interaction term and exhibited significant upregulation in tumor tissues (p < 0.0001, FDR < 0.001). (A) The cluster figure shows tumor tissue expression mean centered to normal tissues for 11 out of the 111 genes examined (log2 (tumor/normal) TPM), color-coded from blue to red, depicting a decrease to an increase in expression in tumor tissues, respectively; they were further filtered using significant hazard ratio calculations for TGFB2 mRNA expression (p < 0.05) AND Gene2 expression (p < 0.05) AND TGFB2 by Gene2 interaction terms (p < 0.05). (B) The table depicts a comparison of mRNA expression values, comparing normal (n = 179) versus tumor samples (n = 786 evaluable samples), ordered by fold increase.
Ijms 26 11580 g001
Figure 2. Prognostic markers for breast cancer patients with improved prognosis for high TGFB2 and Gene2 mRNA expression, assessed using Kaplan–Meier plots. Breast cancer patients were correlated with OS outcomes investigating the impact of mRNA expression of TGFB2 (median cut-off for high and low levels of expression), further stratified into four groups based on gene expression levels of GDAP1 (A); TBL1XR1 (B); RNFT1 (C); HACL1 (D); SLC27A2 (E); NLE1 (F); and TXNDC16 (G) mRNA expression levels (median cut-off for high and low levels) in these patients. The Kaplan–Meier plots show four stratified curves for each of the Gene2s. Six pairwise comparisons were performed between the four groups of patients (p-value adjusted using the Benjamini-Hochberg (BH) correction shown in the table insets). We identified TGFB2/Gene2 combinations (a. TGFB2high/Gene2low; b. TGFB2high/Gene2high; c. TGFB2low/Gene2low; d. TGFB2low/Gene2high) that exhibited significant curve separation between the two arms of the four survival curves (p < 0.05), whereby high levels of TGFB2 and maker gene mRNA expression resulted in the most improved survival outcomes. (H) The table depicts comparisons of mRNA expression values comparing normal (n = 179) versus tumor samples (n = 786 evaluable samples), ordered by fold increase.
Figure 2. Prognostic markers for breast cancer patients with improved prognosis for high TGFB2 and Gene2 mRNA expression, assessed using Kaplan–Meier plots. Breast cancer patients were correlated with OS outcomes investigating the impact of mRNA expression of TGFB2 (median cut-off for high and low levels of expression), further stratified into four groups based on gene expression levels of GDAP1 (A); TBL1XR1 (B); RNFT1 (C); HACL1 (D); SLC27A2 (E); NLE1 (F); and TXNDC16 (G) mRNA expression levels (median cut-off for high and low levels) in these patients. The Kaplan–Meier plots show four stratified curves for each of the Gene2s. Six pairwise comparisons were performed between the four groups of patients (p-value adjusted using the Benjamini-Hochberg (BH) correction shown in the table insets). We identified TGFB2/Gene2 combinations (a. TGFB2high/Gene2low; b. TGFB2high/Gene2high; c. TGFB2low/Gene2low; d. TGFB2low/Gene2high) that exhibited significant curve separation between the two arms of the four survival curves (p < 0.05), whereby high levels of TGFB2 and maker gene mRNA expression resulted in the most improved survival outcomes. (H) The table depicts comparisons of mRNA expression values comparing normal (n = 179) versus tumor samples (n = 786 evaluable samples), ordered by fold increase.
Ijms 26 11580 g002
Figure 3. Protein–protein associations identify potential links of TGFB2 to the TGFB2-dependent Gene2 having prognostic impact in breast cancers. We integrated data from two screens analyzing TGFB2-dependent Gene2 (Figure 1 and Figure 2) to identify linker proteins, represented as white nodes, that connect TGFB2 (indicated by a red box) with the Gene2 (depicted as colored proteins) using the STRING database. The edges represent the confidence levels of each association, which are based on a combination of experimental evidence, including text mining, laboratory experiments, curated databases, co-expression analysis, neighborhood proximity, gene fusion events, and co-occurrence data. Protein associations were defined by interaction edge scores above 0.4, with line thickness indicating score thresholds at 0.4, 0.7, and 0.9. TGFB2 was associated with the network of GPB4, SULF1, ANTXR1, ITGA11, and COL10A1 via linker proteins via second shell interactors: BGN, COL3A1, COL1A2, COL5A2, and THBS2. A separate cluster of associations consisted of GDAP1, HACL, and SLC27A2. The 2 clusters represented Gene2 that exhibited differential impact on the OS derived from the multivariate Cox proportional hazards model, as distinguished by the HR values for Gene2. The cluster consisting of GPC4, SULF1, ANTXR1, ITGA11, and COL10A1 showed increases in the HR, whereas the cluster consisting of HACL, GDAP1, and SLC7A2 showed decreases in the HR. Red box highlights position of the TGFB2 protein.
Figure 3. Protein–protein associations identify potential links of TGFB2 to the TGFB2-dependent Gene2 having prognostic impact in breast cancers. We integrated data from two screens analyzing TGFB2-dependent Gene2 (Figure 1 and Figure 2) to identify linker proteins, represented as white nodes, that connect TGFB2 (indicated by a red box) with the Gene2 (depicted as colored proteins) using the STRING database. The edges represent the confidence levels of each association, which are based on a combination of experimental evidence, including text mining, laboratory experiments, curated databases, co-expression analysis, neighborhood proximity, gene fusion events, and co-occurrence data. Protein associations were defined by interaction edge scores above 0.4, with line thickness indicating score thresholds at 0.4, 0.7, and 0.9. TGFB2 was associated with the network of GPB4, SULF1, ANTXR1, ITGA11, and COL10A1 via linker proteins via second shell interactors: BGN, COL3A1, COL1A2, COL5A2, and THBS2. A separate cluster of associations consisted of GDAP1, HACL, and SLC27A2. The 2 clusters represented Gene2 that exhibited differential impact on the OS derived from the multivariate Cox proportional hazards model, as distinguished by the HR values for Gene2. The cluster consisting of GPC4, SULF1, ANTXR1, ITGA11, and COL10A1 showed increases in the HR, whereas the cluster consisting of HACL, GDAP1, and SLC7A2 showed decreases in the HR. Red box highlights position of the TGFB2 protein.
Ijms 26 11580 g003
Figure 4. Protein–protein associations identifying potential links of TGFB2 to the PAM50 gene signature. We utilized the STRING database to identify the association of TGFB2 with the PAM50 gene signature. The edges illustrate the confidence levels of each association, which are based on a combination of experimental evidence, such as text mining, laboratory experiments, curated databases, co-expression analysis, neighborhood proximity, gene fusion events, and co-occurrence data. TGFB2 was associated with the network of the PAM50 gene signature via MYC and EGFR. Red box highlights the TGFB2 protein.
Figure 4. Protein–protein associations identifying potential links of TGFB2 to the PAM50 gene signature. We utilized the STRING database to identify the association of TGFB2 with the PAM50 gene signature. The edges illustrate the confidence levels of each association, which are based on a combination of experimental evidence, such as text mining, laboratory experiments, curated databases, co-expression analysis, neighborhood proximity, gene fusion events, and co-occurrence data. TGFB2 was associated with the network of the PAM50 gene signature via MYC and EGFR. Red box highlights the TGFB2 protein.
Ijms 26 11580 g004
Figure 5. Protein–protein interactions identify potential correlations between TGFB2 and the OncotypeDx gene signature. We utilized the STRING database to identify the association of TGFB2 with the OncotypeDx gene signature. The edges represent the confidence levels of each association, which are based on a combination of experimental evidence, including text mining, laboratory experiments, curated databases, co-expression analysis, neighborhood proximity, gene fusion events, and co-occurrence data. TGFB2 was associated with the network of the OncotypeDX gene signature via TGFB1 associations with CD68, PGR, ESR1, BCL2, and ERBB2 proteins. Red box highlights the TGFB2 protein.
Figure 5. Protein–protein interactions identify potential correlations between TGFB2 and the OncotypeDx gene signature. We utilized the STRING database to identify the association of TGFB2 with the OncotypeDx gene signature. The edges represent the confidence levels of each association, which are based on a combination of experimental evidence, including text mining, laboratory experiments, curated databases, co-expression analysis, neighborhood proximity, gene fusion events, and co-occurrence data. TGFB2 was associated with the network of the OncotypeDX gene signature via TGFB1 associations with CD68, PGR, ESR1, BCL2, and ERBB2 proteins. Red box highlights the TGFB2 protein.
Ijms 26 11580 g005
Figure 6. Gene expression levels for genes that showed significant prognostic impact for both the TCGA and KMplotter datasets. (A) The bar charts depict the expression values (mean ± SEM) of 6 significantly upregulated genes in ovarian cancer tumor tissues. They were cross-referenced in the TCGA and KMplotter datasets and ordered according to descending mRNA expression levels in the normal tissue. (B) The table depicts a comparison of mRNA expression values comparing normal (n = 179) versus tumor samples (n = 786 evaluable samples), ordered by expression levels in tumor tissue.
Figure 6. Gene expression levels for genes that showed significant prognostic impact for both the TCGA and KMplotter datasets. (A) The bar charts depict the expression values (mean ± SEM) of 6 significantly upregulated genes in ovarian cancer tumor tissues. They were cross-referenced in the TCGA and KMplotter datasets and ordered according to descending mRNA expression levels in the normal tissue. (B) The table depicts a comparison of mRNA expression values comparing normal (n = 179) versus tumor samples (n = 786 evaluable samples), ordered by expression levels in tumor tissue.
Ijms 26 11580 g006
Figure 7. Negative prognostic impacts of Gene2 mRNA levels that were significantly upregulated in tumor tissue identified in the TCGA cohort of patients. The parameters were calculated from the Cox proportional hazards model to assess the individual effects of ENO1 (A), GLRX2 (B), PLOD1 (C), PRDX4 (D), TAGLN2 (E), and TMED9 (F) mRNA expression levels controlling for TGFB2 mRNA levels, age at diagnosis, treatment (chemo-only versus all other treatments), breast cancer subtype, and the interaction between TGFB2 and Gene2. These parameters were utilized to generate predictive OS curves (mean age = 58 yrs, n = 789, # death events = 59 for chemo-only treated patients, mean TGFB2 mRNA levels, HER2+ patients) for genes exhibiting negative prognostic impact in the TCGA cohort of patients. The OS curves depict the proportion of patients surviving up to 120 months for increasing Zscores (minimum level of mRNA expression (Min), lower quartile for the range of mRNA expression (LQ Range), range of mRNA expression (Med), upper quartile for the range of mRNA expression (UQ Range) and maximum expression (Max)) of each gene (Table 1). Hazard ratios with 95% confidence intervals show a pronounced negative prognostic impact across gene expression levels. The highest level of TAGLN2 mRNA expression exhibited the shortest median OS times (14.7 months).
Figure 7. Negative prognostic impacts of Gene2 mRNA levels that were significantly upregulated in tumor tissue identified in the TCGA cohort of patients. The parameters were calculated from the Cox proportional hazards model to assess the individual effects of ENO1 (A), GLRX2 (B), PLOD1 (C), PRDX4 (D), TAGLN2 (E), and TMED9 (F) mRNA expression levels controlling for TGFB2 mRNA levels, age at diagnosis, treatment (chemo-only versus all other treatments), breast cancer subtype, and the interaction between TGFB2 and Gene2. These parameters were utilized to generate predictive OS curves (mean age = 58 yrs, n = 789, # death events = 59 for chemo-only treated patients, mean TGFB2 mRNA levels, HER2+ patients) for genes exhibiting negative prognostic impact in the TCGA cohort of patients. The OS curves depict the proportion of patients surviving up to 120 months for increasing Zscores (minimum level of mRNA expression (Min), lower quartile for the range of mRNA expression (LQ Range), range of mRNA expression (Med), upper quartile for the range of mRNA expression (UQ Range) and maximum expression (Max)) of each gene (Table 1). Hazard ratios with 95% confidence intervals show a pronounced negative prognostic impact across gene expression levels. The highest level of TAGLN2 mRNA expression exhibited the shortest median OS times (14.7 months).
Ijms 26 11580 g007
Figure 8. Validation of TCGA prognostic biomarkers using the independent KMplotter dataset for breast cancer patients. Six genes also significantly impacted OS when the KMplotter web tool was used to access breast cancer data from the Affymetrix dataset (n = 1879) and compare high versus low expression using the median cut-off for the best-selected JetSet probe set for ENO1/ENO1L1 (A), GLRX2 (B), PLOD1 (C), PRDX4 (D), TAGLN2 (E), and TMED9 (F).
Figure 8. Validation of TCGA prognostic biomarkers using the independent KMplotter dataset for breast cancer patients. Six genes also significantly impacted OS when the KMplotter web tool was used to access breast cancer data from the Affymetrix dataset (n = 1879) and compare high versus low expression using the median cut-off for the best-selected JetSet probe set for ENO1/ENO1L1 (A), GLRX2 (B), PLOD1 (C), PRDX4 (D), TAGLN2 (E), and TMED9 (F).
Ijms 26 11580 g008
Table 1. Impact on OS from increasing the mRNA levels of the negative prognostic Gene2 markers.
Table 1. Impact on OS from increasing the mRNA levels of the negative prognostic Gene2 markers.
mRNAZscoremOS (months)
ENO1−1.08>120
ENO11.27>120
ENO13.62>120
ENO1693.8
ENO18.3254.2
GLRX2−1.36>120
GLRX21.56>120
GLRX24.47114.8
GLRX27.3961.9
GLRX210.331
PLOD1−0.92>120
PLOD11.41>120
PLOD13.73114.8
PLOD16.0661.94
PLOD18.3932.09
PRDX4−1.16>120
PRDX41.03>120
PRDX43.22>120
PRDX45.4295.7
PRDX47.6162.5
TAGLN2−1.54>120
TAGLN20.92>120
TAGLN23.3793.8
TAGLN25.8332.6
TAGLN28.2814.7
TMED9−1.72>120
TMED90.48>120
TMED92.67>120
TMED94.8695.7
TMED97.0662.5
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

Qazi, S.; Richardson, S.; Potts, M.; Myers, S.; Trieu, V. Bioinformatic Approach to Identify Positive Prognostic TGFB2-Dependent and Negative Prognostic TGFB2-Independent Biomarkers for Breast Cancers. Int. J. Mol. Sci. 2025, 26, 11580. https://doi.org/10.3390/ijms262311580

AMA Style

Qazi S, Richardson S, Potts M, Myers S, Trieu V. Bioinformatic Approach to Identify Positive Prognostic TGFB2-Dependent and Negative Prognostic TGFB2-Independent Biomarkers for Breast Cancers. International Journal of Molecular Sciences. 2025; 26(23):11580. https://doi.org/10.3390/ijms262311580

Chicago/Turabian Style

Qazi, Sanjive, Stephen Richardson, Mike Potts, Scott Myers, and Vuong Trieu. 2025. "Bioinformatic Approach to Identify Positive Prognostic TGFB2-Dependent and Negative Prognostic TGFB2-Independent Biomarkers for Breast Cancers" International Journal of Molecular Sciences 26, no. 23: 11580. https://doi.org/10.3390/ijms262311580

APA Style

Qazi, S., Richardson, S., Potts, M., Myers, S., & Trieu, V. (2025). Bioinformatic Approach to Identify Positive Prognostic TGFB2-Dependent and Negative Prognostic TGFB2-Independent Biomarkers for Breast Cancers. International Journal of Molecular Sciences, 26(23), 11580. https://doi.org/10.3390/ijms262311580

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop