Open Access

Revolutionizing prognostic predictions in colorectal cancer: Macrophage‑driven transcriptional insights from single‑cell RNA sequencing and gene co‑expression network analysis

  • Authors:
    • Yang Feng
    • Zhuo Cheng
    • Jingyuan Gao
    • Tao Huang
    • Jun Wang
    • Qian Tang
    • Ke Pu
    • Chang Liu
  • View Affiliations

  • Published online on: October 3, 2024     https://doi.org/10.3892/ol.2024.14721
  • Article Number: 587
  • Copyright: © Feng et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Tumor‑associated macrophages have become important biomarkers for cancer diagnosis, prognosis and therapy. The dynamic changes in macrophage subpopulations significantly impact the outcomes of cancer immunotherapy. Hence, identifying additional macrophage‑related biomarkers is essential for enhancing prognostic predictions in colorectal cancer (CRC) immunotherapy. CRC single‑cell RNA sequencing (scRNA‑seq) data was obtained from the Gene Expression Omnibus (GEO) database. The data were processed, normalized and clustered using the ‘Seurat’ package. Cell types within each cluster were annotated using the ‘SingleR’ package. Weighted gene co‑expression network analysis identified modules corresponding to specific cell types. A non‑negative matrix factorization algorithm was employed to segregate different clusters based on the selected module. Differentially expressed genes (DEGs) were identified across various clusters and a prognostic model was constructed using lasso regression and Cox regression analyses. The robustness of the model was validated using The Cancer Genome Atlas (TCGA) database and GEO microarrays. Additionally, the prognosis, immune characteristics and response to immune checkpoint inhibitor (ICI) therapy were individually analyzed. The scRNA‑seq data from GSE200997, consisting of 23 samples, were analyzed. Dimensionality reduction and cluster identification allowed the isolation of the primary myeloid cell subpopulations. The macrophage‑related brown module was identified, which was further divided into two clusters. Using the DEGs from these clusters, a prognostic model was developed, comprising five macrophage‑related genes. The robustness of the model was confirmed using microarray datasets GSE17536, GSE38832 and GSE39582, as well as TCGA cohort. Patients classified as high‑risk by the present model exhibited poorer survival rates, lower tumor mutation burden, reduced microsatellite instability, lower tumor purity, more severe tumor immune dysfunction and exclusion, and less benefit from ICIs therapy compared with low‑risk patients. The present prognostic model shows promise as a biomarker for risk stratification and predicting therapeutic efficacy in patients with CRC. However, further well‑designed prospective studies are necessary to validate the findings.

Introduction

Colorectal cancer (CRC) represents a significant global health burden, contributing to a substantial number of cancer-related deaths worldwide. In 2020 alone, there were over 1.9 million new cases and 0.94 million deaths attributed to CRC globally (1). It ranks third in terms of incidence and second in mortality among cancers (1). Despite the potential of early screening to reduce CRC incidence and mortality, challenges persist in the performance of screening tests and patient adherence among eligible populations (2). Metastatic CRC occurs in ~20% of all CRC cases (3), with 40% experiencing recurrence following treatment of the primary lesion (4). Unfortunately, the prognosis for metastatic CRC remains grim, with a five-year survival rate of <20% (5). Elderly individuals bear a significant burden, as nearly 70% of CRC cases are diagnosed in those aged >65 years (6). However, the impact of age on survival outcomes is not universally agreed upon. Factors such as stage at presentation, tumor location, preexisting health conditions and treatment type may confound the prognosis of older patients (7). Given these challenges, there is an urgent need to develop more effective therapeutic strategies for patients with advanced-stage CRC.

Cancer immunotherapy has emerged as a promising approach for treating challenging solid tumors by enhancing the ability of the immune system to eliminate cancer cells (810). In CRC, immune checkpoint inhibitors (ICIs) gained regulatory approval following the CheckMate142 clinical trial in 2017 (11,12). The ICIs are specifically designed for patients with CRC with a high tumor mutation burden (TMB), characterized by deficient mismatch repair protein (dMMR) and high microsatellite instability (MSI-H), collectively known as the dMMR/MSI-H CRC subtype (12). PD-L1 expression in immune cells is significantly higher in MSI-H CRC than in proficient MMR [low microsatellite instability (MSI-L)] tumors, with no notable differences among various MSI-H molecular subtypes (13). Currently, the recommended screening for defective MMR involves immunohistochemistry (IHC) and/or MSI testing (14). However, capturing the biological and technical heterogeneity of MSI testing poses challenges. IHC testing of the mismatch repair machinery may yield varying results for specific germline mutations, while somatic nonsense mutations can also influence the overall findings (14). Consequently, it is crucial for CRC immunotherapy to identify the molecular characteristics of the tumor microenvironment (TME) and search for reliable immune prognostic indicators.

Macrophages are pivotal in various immune processes, serving essential functions such as phagocytosis, antigen presentation and the secretion of signaling molecules (15,16). In the TME, tumor-associated macrophages (TAMs) are derived from peripheral blood monocytes that infiltrate tumor tissues (17,18). These TAMs are closely associated with tumor initiation, progression, angiogenesis and metastasis (19). TAMs are generally classified into two subpopulations including classically activated macrophages (M1) and alternatively activated macrophages (M2) (20). M1 macrophages release chemokines and pro-inflammatory cytokines, which have anti-tumor effects and promote immune surveillance (20). By contrast, M2 macrophages secrete inhibitory cytokines that primarily support tumor growth and metastasis (21).

High levels of TAM infiltration in tumor tissues are typically considered a risk factor for poor prognosis in cancer treatments, including radiotherapy, chemotherapy and targeted therapy (16,2224). Moreover, the dynamic changes in macrophage subpopulations can significantly influence the effectiveness of immunotherapy across various cancer types (2527). Conventional biomarkers are often insufficient for predicting the efficacy of cancer immunotherapies. Although gene mutations such as V-Ki-Ras2 Kirsten rat sarcoma 2 viral oncogene (KRAS), neuroblastoma RAS viral oncology and B-Raf Proto-Oncogene (BRAF), inflammatory markers such as neutrophil-lymphocyte ratio, lymphocyte-monocyte ratio and platelet-lymphocyte ratio, and aberrant miRNAs serve as prognostic and predictive biomarkers for personalized CRC therapy, more research is required to optimize their detection and validation (28). In the present study, to identify more macrophage-related biomarkers with clinical relevance, single-cell RNA sequencing (scRNA-seq) was used for a precise analysis of CRC macrophages.

scRNA-seq is a revolutionary method that enables the detailed examination of global gene expression profiles in individual cell types, providing profound insights into cellular heterogeneity (29,30). Currently, numerous research initiatives aim to discover novel biomarkers for malignancies by integrating scRNA-seq with traditional RNA sequencing (RNA-seq) (3133). Despite these efforts, there remains a significant gap in knowledge regarding macrophage-related immunotherapeutic indicators identified through the combination of scRNA-seq and weighted gene co-expression network analysis (WGCNA) (34). The present study aimed to bridge this gap by developing a novel gene signature through the integration of the aforementioned advanced tools, thereby improving prognostic predictions for CRC immunotherapy. This innovative approach offers a theoretical foundation for creating personalized treatment strategies for patients with CRC.

Materials and methods

Data acquisition and processing

The present study utilized five independent public datasets, including scRNA-seq, high-throughput RNA-seq and microarray cohorts. The scRNA-seq dataset (GSE200997; n=23) (35), was sourced from the Gene Expression Omnibus (GEO) (https://www.ncbi.nlm.nih.gov/geo/). This dataset provided valuable insights into colorectal cellular diversity and heterogeneity within tumor and microenvironmental cells. Transcriptome datasets and corresponding clinical information for CRC were obtained from The Cancer Genome Atlas [dataset no. TCGA-COAD/TCGA-READ; n=522; colon cancer/rectum cancer (481/41); https://portal.gdc.cancer.gov/]. Additionally, microarray cohorts from the GEO database were included: GSE17536 (n=177), GSE38832 (n=122) and GSE39582 (n=585).

The raw RNA-seq count data were converted to transcripts per million format and subsequently log-2 transformed. Expression profiles from GEO were processed and normalized using the ‘affy’ (version 1.48.0) and ‘lumi’ (version 2.22.0) packages, tailored to the different platforms. The four datasets, excluding the scRNA-seq cohort, were combined, and the ‘sva’ package (3.18.0) was employed to correct for batch effects.

scRNA-seq data analysis

ScRNA-seq data analysis was performed with R language programming (R version 4.2.3, http://www.r-project.org). To ensure quality, single-cell gene expression profiles were filtered to remove mitochondrial genes and cells with <200 detected genes. Dimensional reduction and clustering visualization were performed using the ‘Seurat’ package. The resulting cell clusters were annotated using the ‘SingleR’ package. Unique marker genes for each cluster were identified with the ‘FindAllMarkers’ function. Subsequently, enrichment analysis was conducted using the irGSEA package to gain further insights into the biological significance of these markers.

WGCNA

To explore the correlations between gene modules and clinicopathological data, WGCNA (version 1.72–5) was used (36). Initially, a scale-free gene co-expression networks was constructed using the ‘wgcna’ package, removing outlier samples with a connectivity threshold <-2.5. The soft threshold powers were determined by calculating the scale-free topology fitting indices (R2) using the ‘pickSoftThreshold’ function, ensuring a power value >0.8 to approximate a scale-free network topology.

After which, the adjacency matrix was transformed into a topological overlap matrix (TOM) and computed the corresponding dissimilarity TOM (1-TOM). Using the dynamic tree cut method, gene modules were identified and colored. The relationship between module eigenvalues and phenotypes was evaluated, selecting the modules with the highest correlation for further analysis.

Non-negative matrix factorization and estimation of TME cell infiltration

To identify genes associated with prognosis, univariate Cox regression analysis was conducted. Molecular clustering was performed using the non-negative matrix factorization (NMF) package (R version 4.2.3; http://cran.r-project.org/package=NMF) (37), iterating through the matrix factorization process. The ESTIMATE algorithm was applied to infer immune and stromal scores for each sample, providing insights into the TME. Additionally, the CIBERSORT algorithm was utilized to estimate the enrichment scores of immune and stromal cell types within the samples (38).

Construction of the Macrophages-Gene-Hub-Signature-related prognostic model

Using the ‘limma’ package (version 3.28.6), 175 differentially expressed genes (DEGs) were identified between two subtypes. To mitigate the risk of overfitting, a Macrophages-Gene-Hub-Signature (MHGS) risk score was developed using the Lasso regression. Subsequently, multivariate Cox regression analysis was conducted to screen candidate genes. The MHGS score was calculated as follows: MHGS score=Σ (Exp × coefi), where ‘Exp’ and ‘coefi’ denote the expression and coefficient of each MHGS-related gene, respectively. Patients in both the training and validation sets were stratified into low-risk and high-risk groups based on the median MHGS score. Kaplan-Meier survival curves were generated and analyzed for these groups. Additionally, the prognostic accuracy of the risk-score model was assessed using receiver operating characteristic (ROC) curves.

Analysis of molecular and immune characteristics and ICI therapy in the MHGS model

The analysis was initiated by performing differential expression analysis of all genes between high-risk and low-risk MHGS groups using the ‘limma’ package (version 3.28.6). Subsequently, gene set enrichment analysis (GSEA) was employed via the ‘clusterProfiler’ package in R (R version 4.2.3; http://www.r-project.org) to identify specific signaling pathways associated with these genes, focusing on those with statistical significance (P<0.05 and FDR<0.25).

Gene mutation analysis utilized data obtained from the TCGA and GEO databases, leveraging the ‘Maftools’ package (version 0.99.30) to assess genetic alterations across different risk groups. Pearson correlation analysis was conducted to explore the relationship between the MHGS score and TMB. Additionally, the relative proportions of 22 types of immune cells in the MHGS groups were estimated using the CIBERSORT algorithm. These proportions were then compared alongside clinicopathological factors such as age, sex, TNM stage, TP53, KRAS, BRAF and recurrence.

To evaluate the predictive value of the MHGS score in the context of immunotherapy, the Tumor Immune Dysfunction and Exclusion (TIDE) score (http://tide.dfci.harvard.edu) and the immunophenoscore (IPS) were employed from The Cancer Immunome Atlas (TCIA) (https://tcia.at/) to assess treatment response. A lower TIDE score and higher IPS are indicative of a more favorable response to immunotherapy (39). Survival analyses of the MHGS risk score were conducted within a cohort of patients with urothelial cancer treated with anti-PD-L1 therapy (40). Furthermore, time-dependent ROC curve analyses were performed to compare the prognostic value of MHGS with that of the tumor inflammation signature (TIS), calculated as the average log2-scale normalized expression of 18 signature genes (41).

Cell culture and transfection procedures

CRC cell lines (RKO, SW480 and LoVo), colon epithelial cells (HIEC) and 293T cells were cultured in high-glucose DMEM medium (HyClone; Cytiva) supplemented with 10% fetal bovine serum (FBS; Gibco; Thermo Fisher Scientific, Inc.). HCT15 cells were cultured in RPMI 1640 medium (HyClone; Cytiva) with 10% FBS (Gibco; Thermo Fisher Scientific, Inc.). These cell lines were sourced from the Sichuan Bio Biotechnology Co., Ltd. and maintained at 37°C with 5% CO2 in culture dishes.

To downregulate secreted protein acidic and rich in cysteine-like 1 (SPARCL1) expression, shRNA targeting SPARCL1 was obtained from Shanghai GenePharma Co., Ltd. and transfected using Lipofectamine® 2000 (Invitrogen; Thermo Fisher Scientific, Inc.) following the manufacturer's protocol. The specific siRNA sequences used were: sh-SPARCL1 target sequence: CCGGCCCGACAAATGCAAGATTATTCTCGAGAATAATCTTGCATTTGTCGGGTTT; sh-negative control (sh-NC) target sequence: CCTAAGGTTAAGTCGCCCTCGCTCGAGCGAGGGCGACTTAACCTTAGG. PLKO.1-TRC was selected as the plasmid backbone. The concentration of nucleic acid was 2 µg plasmid DNA per well in a 6-well plate. The transfection was carried out at 37°C for 4–6 h. A 48-h interval was maintained between transfection and subsequent experimentation.

Tissue microarray (TMA) and IHC analysis

A total of 80 pairs of CRC and adjacent tissue samples were collected from the Second Affiliated Hospital of Xi'an Jiaotong University (Xi'an, China) during the period from January 2023 to September 2023. The ages of the patients ranged from 31 to 81 years, comprising 36 women and 44 men. The inclusion criteria included: i) Aged over 18 years; ii) diagnosis of advanced-stage colorectal cancer based on pathological standards; iii) availability of complete clinicopathological information including sex, age, TNM stage, treatment and prognosis; iv) willingness to participate in the clinical study and provision of informed consent; and v) well-preserved tumor tissue sample meeting experimental requirements. The exclusion criteria included: i) Presence of simultaneous or metachronous multi-site tumors; ii) incomplete tumor clinic information leading to insufficient data collection; and iii) loss of tissue sample during transport.

The collection of human tissues was approved by The Medical Ethics Committee of The Second Affiliated Hospital of Xi'an Jiaotong University (Xi'an, China; approval no. 2023R063) and written informed consent was obtained from all individuals or individuals' guardians. The samples were then sent to Hunan Aifang Biotechnology Co., Ltd. where TMA construction and IHC staining analysis was performed on the collected samples. Tissues were dehydrated in ethanol solutions (75, 85, 95 and 100%) for 1 h each, cleared in xylene tanks (Tank I for 20 min; Tank II for 30 min), and infiltrated in paraffin tanks (Tank I for 1 h; Tank II for 1.5 h; Tank III for 2 h). After which, the samples were embedded in liquid paraffin, trimmed, marked and retrieved for sectioning. Tissue strips were arranged in paraffin blocks for microarray preparation. The final steps involved securing tissue pieces in paraffin blocks using a specialized machine. During sectioning, the paraffin block was adjusted in a microtome, the section was aligned, cut to 4-µm thick, warmed, affixed to a glass slide, dried briefly, heated and baked.

IHC staining for SPARCL1 was performed using the anti-SPARCL1 antibody following standardized protocols described in previous studies (42). Paraffin-embedded samples were fixed with 10% formalin at room temperature (25°C) for 24 h, resin-embedded in paraffin and sectioned at 4 µm thickness. Antigen retrieval was performed at 100°C using phosphate-buffered saline, followed by rehydration in a descending ethanol series. The samples were blocked with 3% BSA at room temperature (25°C) for 30 min (Wuhan Servicebio Technology Co., Ltd.; cat. no. G5001), and 3% hydrogen peroxide was used to block endogenous peroxidase activity in HRP/DAB staining. The samples were incubated with the primary antibody for SPARCL1 (cat. no. 13517-1-AP; Proteintech Group, Inc.; 1:100) at 4°C overnight, followed by incubation with the secondary antibody (polymer-horseradish peroxidase conjugated goat anti-rabbit Ig G polyclonal antibody; cat. no. AFIHC003; Hunan Aifang Biotechnology Co., Ltd) at room temperature for 50 min. DAB was used for chromogen detection, and hematoxylin counterstaining was performed at room temperature for 3 min. Digital images of the stained CRC tissues were captured using a KF-FL-020 digital slide scanner (KonFoong Bioinformation Tech Co., Ltd.) utilizing 50 and 200 µm scale bars and analyzed using Visiopharm software (https://visiopharm.com).

An IHC staining score was calculated for each slide to assess SPARCL1 expression. The score was derived by multiplying the staining intensity (i) by the percentage of positively stained cells (pi) and summing these values: IHC-Score=∑(pi × i). Staining intensity was graded on a scale from 0 (no staining) to 3 (strong staining) (43). Thus, the resulting IHC-Score ranged from 0–300, with higher scores indicating greater overall positive staining intensity.

Cell culture and colony formation assay

Colonies were identified as clusters consisting of >50 cells that originated from a single cell. Cell lines were trypsinized to obtain a single-cell suspension, and 200 cells were seeded per well in 6-well plates. The plates were then incubated for 2 weeks under standard conditions in a 37°C, humidified atmosphere with 5% CO2 using culture medium with 10% FBS and antibiotics. After incubation, colonies were fixed with 4% paraformaldehyde for 30 min at room temperature (25°C) and subsequently stained with 0.5% crystal violet for 1 h at 37°C. The colonies were quantified by manual counting under a light microscope after crystal violet staining. Three independent experiments were performed in triplicate, and the average number of colonies was calculated.

Wound healing assay

The wound closure was assessed by measuring the wound area at 0, 24, and 36 h. The percentage closure was determined by comparing the areas at each time point to the initial area at 0 h. CRC cells treated with various conditions (sh-NC and sh-SPARCL1) were seeded into 6-well culture dishes and allowed to grow until they reached 90% confluence. Subsequently, all plates were placed in a 37°C humidified atmosphere with 5% CO2 for the 36-h experimental duration. A linear scratch was made across the cell monolayer in each well using a 200 µl pipette tip. After creating the wound, the wells were washed with phosphate-buffered saline (PBS) to remove any detached cells. Subsequently, fresh medium without FBS was added to the wells. The progress of wound closure was monitored and recorded using an inverted microscope (Olympus Corporation) at 0, 24 and 36 h post-scratching. A total of three independent experiments were conducted in triplicate to calculate the average percentage of wound closure.

Cell migration assay

For the cell migration assay, 3×104 cells were suspended in serum-free medium and seeded into the upper chamber of a Transwell plate (Corning Inc.). Each lower chamber was filled with medium containing 10% FBS (Gibco; Thermo Fisher Scientific, Inc.; 500 µl/well). The cells were allowed to migrate for 24 h at 37°C. Following incubation, non-migrated cells on the upper surface of the membrane were carefully removed using a cotton swab. Subsequently, the migrated cells on the lower surface of the membrane were fixed with 4% paraformaldehyde for 15 min at room temperature (25°C) and the fixed cells were stained with a 0.1% crystal violet solution for 30 min. Finally, the stained cells were quantified by counting by the number of stained cells in five randomly selected fields per membrane under a light microscope.

Protein extraction and Western blot analysis

Total protein was extracted using RIPA lysis buffer (Beyotime Institution of Biotechnology) supplemented with protease and phosphatase inhibitors (Beyotime Institution of Biotechnology). The concentration of extracted proteins was quantified using a BCA protein assay kit (Beyotime Institution of Biotechnology). Subsequently, the proteins were separated by 10% SDS-PAGE and transferred onto a PVDF membrane (Cytiva) with 30 µg protein loaded per lane. After which, 5% non-fat dry milk in TBST (Tris-buffered saline with 0.1% Tween-20) was used to block non-specific antigen sites for 1 h at room temperature (25°C). The membranes were incubated overnight at 4°C with the following primary antibodies: SPARCL1 (1:1,000 dilution; cat. no. ab255597; Abcam) and β-actin (1:1,000 dilution; cat. no. ab8227; Abcam), followed by incubation for 1 h at room temperature (20°C) with horseradish peroxidase-conjugated goat anti-rabbit secondary antibodies (1:50,000, cat. no. HA1001; 1:20,000; cat. no. HA1006; HUABIO) for visualization of protein bands. The transferred proteins were detected using an enhanced chemiluminescence detection system (Sensi Sage Technology).

Statistical analysis

Data are presented as the mean ± standard deviation (SD) for continuous variables. For non-parametric data, results were presented as the median with interquartile range (IQR). Categorical variables were presented as frequencies and percentages. Graphs display individual data points along with error bars representing SD or IQR as appropriate. The statistical analyses were conducted using R (version 3.3.1) and GraphPad Prism 8.0 (Dotmatics). Comparisons between two groups were evaluated by the Wilcoxon rank-sum test, while the Kruskal-Wallis test was conducted to compare more than two groups. An independent t-test was applied to compare continuous variables between two groups. P<0.05 was considered to indicate a statistically significant difference.

Results

Single-cell analysis reveals cell subtypes

scRNA-seq analysis of the GSE200997 dataset included 16 colon cancer samples and seven normal samples, totaling 47,560 immune cells passing quality control (Fig. S1). Using principal component analysis (Figs. S2 and S3) and t-distributed stochastic neighbor embedding (t-SNE) analysis (Fig. 1A and B), the cells were classified into 31 clusters. Based on distinct gene signatures, the immune cells were categorized into seven major clusters: i) T cells; ii) B cells; iii) epithelial cells; iv) natural killer (NK) cells; v) myeloid cells; vi) fibroblasts; and vii) endothelial cells (Fig. 1C and D). Further analysis revealed significant decreases in T cells, B cells and NK cells in tumor tissues, accompanied by increases in epithelial cells and myeloid cells compared with normal tissues (Fig. 1E). Given the pivotal role of macrophages in tumor immunity and progression, additional investigation and characterization of macrophage cells were conducted at the single-cell level in the present study.

Identification of macrophage-related genes in CRC development via WGCNA

The myeloid cell subpopulations underwent t-SNE dimensional reduction, revealing six principal subclusters (Fig. 2A). Analysis of specific gene markers in each subcluster indicated that cluster one showed elevated expression of CD68 and CD14 (Fig. S4), recognized as distinctive macrophage markers. To further investigate these subclusters, co-expression network analysis was conducted using the WGCNA package. A soft thresholding power β of seven was chosen, achieving a fit index of 0.90 and demonstrating a network with scale-free topology (Fig. 2B). Dynamic tree cutting identified ten modules (Figs. S5 and 2C), with the brown, yellow, turquoise and purple modules strongly correlating with clusters 1, 2, 3 and 5, respectively (Fig. 2D). Given the association of cluster 1 with macrophages, the blue module linked to cluster 1 was selected for network analysis. The network connectivity of the top 25 hub genes within the brown module was visualized (Fig. 2E and F).

Different molecular subtype identification

Based on the expression profiles of 25 macrophage-related genes, the NMF clustering algorithm was applied to classify patients with CRC into two distinct subtypes: Subtype C1 and C2 (Figs. 3A and S6). Notably, patients in subtype C1 exhibited significantly improved overall survival (OS) outcomes compared with those in subtype C2 (Fig. 3B). To explore the relationship between these subtypes and 22 human immune cell subsets within CRC samples, correlation analyses were conducted using the CIBERSORT algorithm (Fig. 3C). The findings revealed substantial differences in immune cell infiltration between the two subtypes, except for naïve B cells, CD8+ T cells, memory activated CD4+ T cells, activated NK cells and resting dendritic cells. Specifically, subtype C1 showed higher infiltration levels of memory B cells, plasma cells, naïve CD4+ T cells, memory resting B cells, regulatory T cells, resting NK cells, monocytes, activated dendritic cells and activated mast cells compared with subtype C2. By contrast, the infiltration levels of follicular helper T cells, delta T cells (Tγδ), M0, M1 and M2 macrophages, resting mast cells, eosinophils and neutrophils were significantly lower in subtype C1 than in subtype C2. Additionally, the TME scores were evaluated using the ESTIMATE algorithm, including stromal score, immune score and estimate score, for both subtypes. The analysis revealed that subtype C1 displayed inferior stromal and immune scores but higher tumor purity compared with subtype C2 (Fig. 3D).

MGHS prognostic model construction and validation

Using the ‘limma’ package, an analysis was conducted to identify DEGs associated with CRC subtypes. Initially, 175 DEGs underwent univariate Cox regression to pinpoint potential prognostic markers within the CRC cohorts. To streamline the findings, LASSO regression and Cox regression were used to select independent prognostic markers (Fig. 4A). This comprehensive approach highlighted five genes, V-set and immunoglobulin domain containing 4 (VSIG4), CXCL9, secreted phosphoprotein 1 (SPP1), CXCL13 and SPARCL1, as significant prognostic indicators. The risk model was constructed using the coefficients of these genes, and the risk score was calculated using the following formula: MHGS risk score=(expression level of VSIG4 × 0.21) + (expression level of CXCL9 × −0.09) + (expression level of SPP1 × 0.07) + (expression level of CXCL13 × −0.14) + (expression level of SPARCL1 × 0.16).

Patients were stratified into high- and low-risk groups based on their risk scores, using the median value as the threshold. The survival analysis demonstrated that patients in the high-risk group had significantly poorer OS compared with those in the low-risk group (Fig. 4B).

Furthermore, the risk score exhibited robust predictive performance for OS in both the training and testing sets. In the training set, the area under the curve (AUC) for predicting survival at 1, 2 and 3 years was 0.681, 0.661 and 0.674, respectively (Fig. 4C). Similarly, in the testing set, the AUC for predicting survival at 1, 3 and 5 years was 0.695, 0.646 and 0.648, respectively (Fig. 4C). Detailed survival outcomes for individual patients in the training and testing sets were depicted using risk plots, providing a comprehensive visualization of patient-specific outcomes based on the risk score (Fig. 4D and E).

Mutation landscape and enrichment analysis between high and low-risk MGHS groups

Numerous studies have demonstrated that cancers with a high mutation burden may benefit from immunotherapy due to increased neoantigen presence (8,44). In the present analysis of mutation data from both training and testing cohorts, it was observed that the low-risk group exhibited a higher TMB compared with the high-risk group. Spearman correlation analysis further indicated a negative correlation between the MHGS score and TMB, suggesting that patients in the low-risk group may experience more favorable outcomes with immunotherapy (Fig. 5A and B).

To delve deeper into somatic mutation characteristics, waterfall plots were generated comparing the two MHGS score groups in both the training and testing cohorts. Consistently, both cohorts revealed a higher frequency of mutations among the top 15 ranked genes in the low-risk group compared with the high-risk group (Fig. 5C and D). Subsequently, GSEA was performed to identify pathways significantly enriched between the two risk groups. Genes in the high-risk group were notably enriched in pathways related to extracellular matrix receptor interaction and focal adhesion (Fig. 5E and F). Conversely, genes in the low-risk group exhibited significant enrichment in pathways associated with cell cycle regulation and DNA replication (Fig. 5E and F).

Association between MHGS score, clinicopathological characteristics and immune cell profiling in CRC

The association between the MHGS score and clinicopathological characteristics was investigated. The MHGS score showed significant associations with TNM stage, survival status and disease occurrence (Fig. 6A). Both univariate and multivariate Cox analyses confirmed the MHGS score as an independent prognostic factor for patients with CRC (Figs. 6A, B and S7). Additionally, differences in the proportions of 22 immune cell types between high and low-risk groups were explored using the CIBERSORT algorithm. Fig. 6C illustrates the proportions of tumor-infiltrating immune cells in these groups, while their distributions are detailed in Fig. 6D through boxplots. Compared with the high-risk group, the low-risk group exhibited significantly higher proportions of B naïve cells, plasma cells, CD8+ T cells, activated memory CD4+ T cells, T follicular helper cells, activated NK cells, macrophage M1 and dendritic cells. Notably, patients with CRC with low-risk scores demonstrated higher infiltration of macrophage M1, whereas the high-risk score group showed elevated levels of macrophage M2 (Fig. 6D).

Association between MHGS score, MSI status, TME and immunotherapy response in CRC: The benefits of ICI therapy in high and low-risk MGHS groups

Recent evidence indicates that patients with MSI-H are more responsive to immunotherapy (45). Correlation analyses between MSI status and MHGS revealed that patients with CRC with MSI-L or microsatellite stable (MSS) status had higher risk scores compared with those with MSI-H in both training and testing sets. This highlights the ability of the prognostic model to distinguish microsatellite status in CRC (Fig. 7A).

A high MHGS score was closely associated with elevated stromal scores and lower tumor purity, while no significant difference in immune scores was observed between high and low-risk groups (Fig. 7B). Next, in the present analysis of MHGS score and tumor immune escape in CRC immunotherapy, T-cell dysfunction and exclusion was focused on. The high-risk group exhibited lower expression levels of markers such as merck18, CD8 and IFNγ compared with the low-risk group, indicating severe T-cell dysfunction associated with a high MHGS score (Fig. 7C). Additionally, the high-risk group showed a stronger association with the cancer-associated fibroblast signature and significantly lower levels of cytotoxic T lymphocytes (CTLs). Combining these findings with dysfunction, exclusion and TIDE scores (Fig. 7C), it became evident that patients with high MHGS scores often had poor immune therapy responses due to T-cell dysfunction and exclusion.

To assess the association between MHGS and immune checkpoint blockade (ICB) response, the IPS was used as a predictive scoring system from TCIA database. Patients with CRC in the low-risk group showed more significant benefits from PD-L1 or/and CTLA-4 immunotherapy compared with those in the high-risk group (Fig. 8A). Validation in the urothelial carcinoma immunotherapeutic cohort (IMvigor210) (40) consistently showed that the high-risk group had a less effective response to immunotherapy compared with the low-risk group (Fig. 8B). Survival analysis further demonstrated that patients in the high-risk group had inferior survival outcomes after receiving ICIs compared with those in the low-risk group (Fig. 8C).

The prognostic accuracy of TIDE and TIS scores was compared with the MHGS score in training and testing sets. The MHGS score exhibited superior prognostic value for patients with CRC (Fig. 8D). Additionally, immunotherapeutic efficacy (Fig. 8E) and OS time (Fig. 8F) was analyzed using these scores for cancer patients treated with ICIs. In both sets, the MHGS score demonstrated improved prognostic accuracy for predicting immunotherapeutic efficacy and OS. However, caution is warranted when using the MHGS score as an immunotherapy indicator, given its AUC value below 70%. Further validation in larger immunotherapy cohorts is essential.

IHC validation of risk score model genes

Numerous studies have extensively investigated the association between CXCL9 (46), SPP1 (47) and CXCL13 (48) with CRC. However, the research concerning the association between VSIG4 and SPARCL1 in relation to CRC is relatively limited. Hence, SPARCL1 was selected for detailed investigation as a target gene in the present study. The clinical analysis revealed that SPARCL1 expression was lower in RKO, SW480 and HCT15 cell lines compared with HIEC cells. By contrast, LoVo cells exhibited high expression levels of SPARCL1 compared with HIEC cells (Fig. 9A and Fig. S8A). To validate these findings, IHC was performed on a tissue microarray. The results demonstrated a significant downregulation of SPARCL1 in CRC tissues compared with adjacent normal tissues (Fig. 9B and Fig. S8B). To elucidate the functional role of SPARCL1, stable knockdown experiments in SW480 and HCT15 cells were conducted (Fig. 9C). Subsequently, the impact of SPARCL1 was assessed on cell proliferation using a colony formation assay, revealing a substantial increase in proliferation upon SPARCL1 knockdown in SW480 and HCT15 cells (Fig. 9D). Furthermore, both wound healing and Transwell assays showed a significant enhancement in migration and invasion of CRC cells upon SPARCL1 silencing (Fig. 9E and F).

Discussion

In the present study, a prognostic model for CRC was developed by integrating single-cell RNA-seq and RNA-seq data. Initially, scRNA-seq analysis was conducted to identify distinct cell subpopulations, focusing particularly on myeloid cell clusters. Using t-SNE dimensional reduction, myeloid cells were categorized into six principal subclusters and specific markers highly expressed in macrophages were identified. Subsequently, WGCNA was employed to explore the association between gene modules and these six clusters. This analysis identified macrophage-related module genes and delineated macrophage-related subtypes (C1 and C2) using NMF. In terms of TME infiltrating characteristics, subtype C1 exhibited a higher proportion of most immune cells but a lower content of macrophages compared with subtype C2. However, using the ESTIMATE algorithm, it was found that patients with the subtype C1 had lower immune and stromal scores but higher tumor purity compared with patients with the C2 subtype. Survival analysis indicated a superior survival advantage for patients with the C1 subtype, potentially linked to their higher somatic mutation burden due to elevated tumor purity (49). Following this, a prognostic model was conducted using DEGs between subtype C1 and C2.

The model identified five genes significantly correlated with CRC prognosis. In the high-risk group, VSIG4, SPP1 and SPARCL1 showed upregulated expression, whereas CXCL9 and CXCL13 were downregulated compared with the low-risk group. VSIG4, located on the X chromosome, has immunosuppressive functions in macrophages, including complement system inhibition and T cell suppression, suggesting its potential as a diagnostic and prognostic biomarker in cancer contexts (5055). SPP1, a member of the SIBLING family, promotes tumorigenesis by activating MMPs and is associated with macrophage M2 polarization and poor prognosis (5660). SPARCL1, belonging to the SPARC family, is implicated in tumor metastasis regulation and prognosis (6163). In the present study, the findings of reduced SPARCL1 protein expression in CRC, supported by western blotting and IHC, associated with increased metastasis risk in functional assays. CXCL9 (64) and CXCL13 (65), members of the chemokine ligand family, play critical roles in antitumor immunity and TME regulation. CXCL9, derived from CD68+ macrophages, confers a survival advantage, whereas CXCL13, associated with M2 macrophages, may promote tumor metastasis (6668). The observed discrepancies between transcriptome and protein levels highlight potential complexities in tumor biology. Despite these challenges, the present model demonstrated moderate accuracy in predicting CRC survival outcomes.

The present study comprehensively analyzed the gene mutation landscape and immune function in both high and low-risk groups of patients with CRC. A total of two distinct cohorts were examined to identify the top 15 mutation genes specific to each risk group, a number of which are well-known driver genes in cancer research (69). Notably, the low-risk group showed a higher frequency of mutations in these genes compared with the high-risk group. The association between the MGHS risk score, MSI, immune cell infiltration patterns and PD-L1 expression was evaluated relevant to immunotherapy. Patients in the high-risk group predominantly exhibited MSS/MSI-L status, characterized by lower CD8+ CTL infiltration and decreased PD-L1 expression levels. By contrast, patients with MSI-H status showed a favorable response to immunotherapy, with reduced risks of recurrence compared with those with MSI-L status (70,71). Recent studies have categorized a specific TME immune type (TMIT I) characterized by abundant infiltration of CD8+ CTLs and high PD-L1 expression, indicating adaptive immune resistance to tumor cells and favorable outcomes with PD-L1/PD-1 immunotherapy (72,73). Conversely, TMIT II type is marked by low infiltration of CD8+ CTLs and minimal PD-L1 expression, suggesting immune indifference towards tumor cells (74). Furthermore, the risk score model was integrated with the TIDE and IPS scoring systems to predict immunotherapy response. TIDE, a computational method based on tumor immune escape characteristics, highlighted differences in gene signatures associated with tumor immune evasion between the risk groups. Although immunosuppressive cell infiltration varied in the high-risk group, T cell exclusion scores remained higher compared with the low-risk group. Notably, the high-risk group exhibited a higher content of TAMM2, which was not observed in the TIDE scoring system. IPS, a superior predictive scoring system for ICB response, indicated that low-risk patients had higher scores and improved responses to immunotherapy, making them more suitable for ICIs treatment.

To validate the predictive ability of the model, MGHS with TIDE and TIS were compared in the urothelial carcinoma immunotherapy cohort. The results demonstrated that MGHS-based immunotherapy predictions were significantly associated with improved prognosis and OS time compared with TIDE and TIS predictions. It is essential to note that while TIDE, IPS and TIS primarily focus on T cell function and status, providing only a partial reflection of the response of the TME to immunotherapy, MGHS consistently demonstrated moderate predictive ability for survival time and immunotherapy prognosis in patients with CRC. Moreover, MGHS comprises only five genes, making it easier to detect compared with TIDE and TIS.

To conclude, in the present study, scRNA-seq, RNA-seq and microarray data were integrated to develop and validate a macrophage-related prognostic model for CRC. A total of two distinct subtypes, C1 and C2, were identified within the CRC population and a prognostic model based on genes differentially expressed in these subtypes was established. The analysis of prognosis and immune characteristics across various risk groups revealed that higher risk scores is associated with poorer survival outcomes, lower tumor mutational burden, MSI-L status, decreased tumor purity and higher TIDE score. The present prognostic model shows promise as a potential biomarker for risk stratification and predicting treatment response in patients with CRC. Future well-designed prospective studies are essential to validate and further explore the clinical implications of the present findings.

Supplementary Material

Supporting Data

Acknowledgements

Not applicable.

Funding

This work was supported by The National Natural Science Foundation of China (grant no. 82204877), Xi'an municipal Health Commission research project (grant no. 2020qn07), The Intramural Fund of North Sichuan Medical College (grant no. CBY21-QD31) and Nanchong City Talent Development Fund (grant no. CBY23-NCR06).

Availability of data and materials

The data generated in the present study may be requested from the corresponding author.

Authors' contributions

YF, QT, CL and KP were involved in the conceptualization, formal analysis, data collection, visualization, and writing the first draft of manuscript; KP, JYG, TH, JW and ZC performed the data validation, statistical analysis, data interpretation and figure editing; and KP, CL and YF supervised the study and modified the final manuscript. All authors read and approved the final version of the manuscript. CL and KP confirm the authenticity of all the raw data.

Ethics approval and consent to participate

The collection of human tissues was approved by The Medical Ethics Committee of the Second Affiliated Hospital of Xi'an Jiaotong University (Xi'an, China; approval no. 2023R063). All methods were carried out in accordance with relevant guidelines and regulations. Written informed consent was obtained from all individuals or individuals' guardians. The study was performed under the principles of the Declaration of Helsinki.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Glossary

Abbreviations

Abbreviations:

TAMs

tumor-associated macrophages

CRC

colorectal cancer

ScRNA-seq

single-cell RNA sequencing

GEO

Gene Expression Omnibus

WGCNA

weighted gene correlation network analysis

NMF

non-negative matrix factorization

TCGA

The Cancer Genome Atlas

TMB

tumor mutation burden

MSI

microsatellite instability

ICIs

immune checkpoint inhibitors

dMMR

deficient mismatch repair protein

MSI-H

high microsatellite instability

pMMR

proficient mismatch repair

MSS

microsatellite stability

TME

tumor microenvironment

TOM

topological overlap matrix

DEGs

differential expression genes

TMA

tissue microarray

IHC

Immunohistochemistry

TIDE

Tumor Immune Dysfunction and Exclusion

IPS

immunophenoscore

TCIA

The Cancer Immunome Atlas

TIS

tumor inflammation signature

GSEA

gene set enrichment analysis

MGHS

macrophages gene hub signature

ICB

immune checkpoint blockade

TMIT

tumor microenvironment immune type

References

1 

Siegel RL, Miller KD, Fuchs HE and Jemal A: Cancer statistics, 2022. CA Cancer J Clin. 72:7–33. 2022. View Article : Google Scholar : PubMed/NCBI

2 

Ladabaum U, Dominitz JA, Kahi C and Schoen RE: Strategies for colorectal cancer screening. Gastroenterology. 158:418–432. 2020. View Article : Google Scholar : PubMed/NCBI

3 

Han L, Dai W, Mo S, Xiang W, Li Q, Xu Y, Cai G and Wang R: Nomogram of conditional survival probability of long-term survival for metastatic colorectal cancer: A real-world data retrospective cohort study from SEER database. Int J Surg. 92:1060132021. View Article : Google Scholar : PubMed/NCBI

4 

Kahi CJ, Boland CR, Dominitz JA, Giardiello FM, Johnson DA, Kaltenbach T, Lieberman D, Levin TR, Robertson DJ and Rex DK: Colonoscopy surveillance after colorectal cancer resection: Recommendations of the US multi-society task force on colorectal cancer. Am J Gastroenterol. 111:337–346; quiz 347. 2016. View Article : Google Scholar : PubMed/NCBI

5 

Biller LH and Schrag D: Diagnosis and treatment of metastatic colorectal cancer: A review. JAMA. 325:669–685. 2021. View Article : Google Scholar : PubMed/NCBI

6 

Siegel RL, Miller KD, Sauer AG, Fedewa SA, Butterly LF, Anderson JC, Cercek A, Smith RA and Jemal A: Colorectal cancer statistics, 2020. CA Cancer J Clin. 70:145–164. 2020. View Article : Google Scholar : PubMed/NCBI

7 

Osseis M, Nehmeh WA, Rassy N, Derienne J, Noun R, Salloum C, Rassy E, Boussios S and Azoulay D: Surgery for T4 colorectal cancer in older patients: Determinants of outcomes. J Pers Med. 12:15342022. View Article : Google Scholar : PubMed/NCBI

8 

Chan TA, Yarchoan M, Jaffee E, Swanton C, Quezada SA, Stenzinger A and Peters S: Development of tumor mutation burden as an immunotherapy biomarker: Utility for the oncology clinic. Ann Oncol. 30:44–56. 2019. View Article : Google Scholar : PubMed/NCBI

9 

Samstein RM, Lee CH, Shoushtari AN, Hellmann MD, Shen R, Janjigian YY, Barron DA, Zehir A, Jordan EJ, Omuro A, et al: Tumor mutational load predicts survival after immunotherapy across multiple cancer types. Nat Genet. 51:202–206. 2019. View Article : Google Scholar : PubMed/NCBI

10 

Riley RS, June CH, Langer R and Mitchell MJ: Delivery technologies for cancer immunotherapy. Nat Rev Drug Discov. 18:175–196. 2019. View Article : Google Scholar : PubMed/NCBI

11 

Overman MJ, Lonardi S, Wong KYM, Lenz HJ, Gelsomino F, Aglietta M, Morse MA, Van Cutsem E, McDermott R, Hill A, et al: Durable clinical benefit with nivolumab plus ipilimumab in DNA mismatch repair-deficient/microsatellite instability-high metastatic colorectal cancer. J Clin Oncol. 36:773–779. 2018. View Article : Google Scholar : PubMed/NCBI

12 

Overman MJ, McDermott R, Leach JL, Lonardi S, Lenz HJ, Morse MA, Desai J, Hill A, Axelson M, Moss RA, et al: Nivolumab in patients with metastatic DNA mismatch repair-deficient or microsatellite instability-high colorectal cancer (CheckMate 142): An open-label, multicentre, phase 2 study. Lancet Oncol. 18:1182–1191. 2017. View Article : Google Scholar : PubMed/NCBI

13 

Ganesh K, Stadler ZK, Cercek A, Mendelsohn RB, Shia J, Segal NH and Diaz LA Jr: Immunotherapy in colorectal cancer: Rationale, challenges and potential. Nat Rev Gastroenterol Hepatol. 16:361–375. 2019. View Article : Google Scholar : PubMed/NCBI

14 

Adeleke S, Haslam A, Choy A, Diaz-Cano S, Galante JR, Mikropoulos C and Boussios S: Microsatellite instability testing in colorectal patients with Lynch syndrome: Lessons learned from a case report and how to avoid such pitfalls. Per Med. 19:277–286. 2022. View Article : Google Scholar : PubMed/NCBI

15 

Krenkel O and Tacke F: Liver macrophages in tissue homeostasis and disease. Nat Rev Immunol. 17:306–321. 2017. View Article : Google Scholar : PubMed/NCBI

16 

Ruffell B and Coussens LM: Macrophages and therapeutic resistance in cancer. Cancer Cell. 27:462–472. 2015. View Article : Google Scholar : PubMed/NCBI

17 

Mantovani A, Bottazzi B, Colotta F, Sozzani S and Ruco L: The origin and function of tumor-associated macrophages. Immunol Today. 13:265–270. 1992. View Article : Google Scholar : PubMed/NCBI

18 

Ginhoux F, Schultze JL, Murray PJ, Ochando J and Biswas SK: New insights into the multidimensional concept of macrophage ontogeny, activation and function. Nat Immunol. 17:34–40. 2016. View Article : Google Scholar : PubMed/NCBI

19 

Mantovani A, Marchesi F, Malesci A, Laghi L and Allavena P: Tumour-associated macrophages as treatment targets in oncology. Nat Rev Clin Oncol. 14:399–416. 2017. View Article : Google Scholar : PubMed/NCBI

20 

Pan Y, Yu Y, Wang X and Zhang T: Tumor-associated macrophages in tumor immunity. Front Immunol. 11:5830842020. View Article : Google Scholar : PubMed/NCBI

21 

An Y and Yang Q: Tumor-associated macrophage-targeted therapeutics in ovarian cancer. Int J Cancer. 149:21–30. 2021. View Article : Google Scholar : PubMed/NCBI

22 

Jung KY, Cho SW, Kim YA, Kim D, Oh BC, Park DJ and Park YJ: Cancers with higher density of tumor-associated macrophages were associated with poor survival rates. J Pathol Transl Med. 49:318–324. 2015. View Article : Google Scholar : PubMed/NCBI

23 

Yeung OW, Lo CM, Ling CC, Qi X, Geng W, Li CX, Ng KT, Forbes SJ, Guan XY, Poon RTP, et al: Alternatively activated (M2) macrophages promote tumour growth and invasiveness in hepatocellular carcinoma. J Hepatol. 62:607–616. 2015. View Article : Google Scholar : PubMed/NCBI

24 

Zhou SL, Zhou ZJ, Hu ZQ, Huang XW, Wang Z, Chen EB, Fan J, Cao Y, Dai Z and Zhou J: Tumor-associated neutrophils recruit macrophages and T-regulatory cells to promote progression of hepatocellular carcinoma and resistance to sorafenib. Gastroenterology. 150:1646–1658.e1617. 2016. View Article : Google Scholar : PubMed/NCBI

25 

Gubin MM, Esaulova E, Ward JP, Malkova ON, Runci D, Wong P, Noguchi T, Arthur CD, Meng W, Alspach E, et al: High-dimensional analysis delineates myeloid and lymphoid compartment remodeling during successful immune-checkpoint cancer therapy. Cell. 175:1014–1030.e1019. 2018. View Article : Google Scholar : PubMed/NCBI

26 

Molgora M, Esaulova E, Vermi W, Hou J, Chen Y, Luo J, Brioschi S, Bugatti M, Omodei AS, Ricci B, et al: TREM2 modulation remodels the tumor myeloid landscape enhancing Anti-PD-1 immunotherapy. Cell. 182:886–900.e817. 2020. View Article : Google Scholar : PubMed/NCBI

27 

Xiong H, Mittman S, Rodriguez R, Moskalenko M, Pacheco-Sanchez P, Yang Y, Nickles D and Cubas R: Anti-PD-L1 treatment results in functional remodeling of the macrophage compartment. Cancer Res. 79:1493–1506. 2019. View Article : Google Scholar : PubMed/NCBI

28 

Boussios S, Ozturk MA, Moschetta M, Karathanasi A, Zakynthinakis-Kyriakou N, Katsanos KH, Christodoulou DK and Pavlidis N: The developing story of predictive biomarkers in colorectal cancer. J Pers Med. 9:122019. View Article : Google Scholar : PubMed/NCBI

29 

Chen H, Ye F and Guo G: Revolutionizing immunology with single-cell RNA sequencing. Cell Mol Immunol. 16:242–249. 2019. View Article : Google Scholar : PubMed/NCBI

30 

Liang L, Yu J, Li J, Li N, Liu J, Xiu L, Zeng J, Wang T and Wu L: Integration of scRNA-Seq and bulk RNA-Seq to analyse the heterogeneity of ovarian cancer immune cells and establish a molecular risk model. Front Oncol. 11:7110202021. View Article : Google Scholar : PubMed/NCBI

31 

Bao X, Shi R, Zhao T, Wang Y, Anastasov N, Rosemann M and Fang W: Integrated analysis of single-cell RNA-seq and bulk RNA-seq unravels tumour heterogeneity plus M2-like tumour-associated macrophage infiltration and aggressiveness in TNBC. Cancer Immunol Immunother. 70:189–202. 2021. View Article : Google Scholar : PubMed/NCBI

32 

Chen Z, Yu M, Yan J, Guo L, Zhang B, Liu S, Lei J, Zhang W, Zhou B, Gao J, et al: PNOC expressed by B cells in cholangiocarcinoma was survival related and LAIR2 could be a T cell exhaustion biomarker in tumor microenvironment: Characterization of immune microenvironment combining single-cell and bulk sequencing technology. Front Immunol. 12:6472092021. View Article : Google Scholar : PubMed/NCBI

33 

Li X, Liao Z, Deng Z, Chen N and Zhao L: Combining bulk and single-cell RNA-sequencing data to reveal gene expression pattern of chondrocytes in the osteoarthritic knee. Bioengineered. 12:997–1007. 2021. View Article : Google Scholar : PubMed/NCBI

34 

Shi L, Mao H and Ma J: Integrated analysis of tumor-associated macrophages and M2 macrophages in CRC: Unraveling molecular heterogeneity and developing a novel risk signature. BMC Med Genomics. 17:1452024. View Article : Google Scholar : PubMed/NCBI

35 

Khaliq AM, Erdogan C, Kurt Z, Turgut SS, Grunvald MW, Rand T, Khare S, Borgia JA, Hayden DM, Pappas SG, et al: Refining colorectal cancer classification and clinical stratification through a single-cell atlas. Genome Biol. 23:1132022. View Article : Google Scholar : PubMed/NCBI

36 

Langfelder P and Horvath S: WGCNA: An R package for weighted correlation network analysis. BMC Bioinformatics. 9:5592008. View Article : Google Scholar : PubMed/NCBI

37 

Brunet JP, Tamayo P, Golub TR and Mesirov JP: Metagenes and molecular pattern discovery using matrix factorization. Proc Natl Acad Sci USA. 101:4164–4169. 2004. View Article : Google Scholar : PubMed/NCBI

38 

Newman AM, Liu CL, Green MR, Gentles AJ, Feng W, Xu Y, Hoang CD, Diehn M and Alizadeh AA: Robust enumeration of cell subsets from tissue expression profiles. Nat Methods. 12:453–457. 2015. View Article : Google Scholar : PubMed/NCBI

39 

Charoentong P, Finotello F, Angelova M, Mayer C, Efremova M, Rieder D, Hackl H and Trajanoski Z: Pan-cancer immunogenomic analyses reveal genotype-immunophenotype relationships and predictors of response to checkpoint blockade. Cell Rep. 18:248–262. 2017. View Article : Google Scholar : PubMed/NCBI

40 

Mariathasan S, Turley SJ, Nickles D, Castiglioni A, Yuen K, Wang Y, Kadel EE III, Koeppen H, Astarita JL, Cubas R, et al: TGFβ attenuates tumour response to PD-L1 blockade by contributing to exclusion of T cells. Nature. 554:544–548. 2018. View Article : Google Scholar : PubMed/NCBI

41 

Ayers M, Lunceford J, Nebozhyn M, Murphy E, Loboda A, Kaufman DR, Albright A, Cheng JD, Kang SP, Shankaran V, et al: IFN-γ-related mRNA profile predicts clinical response to PD-1 blockade. J Clin Invest. 127:2930–2940. 2017. View Article : Google Scholar : PubMed/NCBI

42 

Jiang L, Liu JY, Shi Y, Tang B, He T, Liu JJ, Fan JY, Wu B, Xu XH, Zhao YL, et al: MTMR2 promotes invasion and metastasis of gastric cancer via inactivating IFNγ/STAT1 signaling. J Exp Clin Cancer Res. 38:2062019. View Article : Google Scholar : PubMed/NCBI

43 

Reiner A, Spona J, Reiner G, Schemper M, Kolb R, Kwasny W, Függer R, Jakesz R and Holzner JH: Estrogen receptor analysis on biopsies and fine-needle aspirates from human breast carcinoma. Correlation of biochemical and immunohistochemical methods using monoclonal antireceptor antibodies. Am J Pathol. 125:443–449. 1986.PubMed/NCBI

44 

Snyder A, Makarov V, Merghoub T, Yuan J, Zaretsky JM, Desrichard A, Walsh LA, Postow MA, Wong P, Ho TS, et al: Genetic basis for clinical response to CTLA-4 blockade in melanoma. N Engl J Med. 371:2189–2199. 2014. View Article : Google Scholar : PubMed/NCBI

45 

Zhou KI, Hanks BA and Strickler JH: Management of microsatellite instability high (MSI-H) gastroesophageal adenocarcinoma. J Gastrointest Cancer. 55:483–496. 2024. View Article : Google Scholar : PubMed/NCBI

46 

Wu Z, Huang X, Han X, Li Z, Zhu Q, Yan J, Yu S, Jin Z, Wang Z, Zheng Q and Wang Y: The chemokine CXCL9 expression is associated with better prognosis for colorectal carcinoma patients. Biomed Pharmacother. 78:8–13. 2016. View Article : Google Scholar : PubMed/NCBI

47 

Sathe A, Mason K, Grimes SM, Zhou Z, Lau BT, Bai X, Su A, Tan X, Lee H, Suarez CJ, et al: Colorectal cancer metastases in the liver establish immunosuppressive spatial networking between tumor-associated SPP1+ macrophages and fibroblasts. Clin Cancer Res. 29:244–260. 2023. View Article : Google Scholar : PubMed/NCBI

48 

Cabrero-de Las Heras S, Hernández-Yagüe X, González A, Losa F, Soler G, Bugés C, Baraibar I, Esteve A, Pardo-Cea MÁ, Ree AH, et al: Changes in serum CXCL13 levels are associated with outcomes of colorectal cancer patients undergoing first-line oxaliplatin-based treatment. Biomed Pharmacother. 176:1168572024. View Article : Google Scholar : PubMed/NCBI

49 

Ding L, Getz G, Wheeler DA, Mardis ER, McLellan MD, Cibulskis K, Sougnez C, Greulich H, Muzny DM, Morgan MB, et al: Somatic mutations affect key pathways in lung adenocarcinoma. Nature. 455:1069–1075. 2008. View Article : Google Scholar : PubMed/NCBI

50 

Langnaese K, Colleaux L, Kloos DU, Fontes M and Wieacker P: Cloning of Z39Ig, a novel gene with immunoglobulin-like domains located on human chromosome X. Biochim Biophys Acta. 1492:522–525. 2000. View Article : Google Scholar : PubMed/NCBI

51 

Small AG, Al-Baghdadi M, Quach A, Hii C and Ferrante A: Complement receptor immunoglobulin: A control point in infection and immunity, inflammation and cancer. Swiss Med Wkly. 146:w143012016.PubMed/NCBI

52 

Helmy KY, Katschke KJ Jr, Gorgani NN, Kljavin NM, Elliott JM, Diehl L, Scales SJ, Ghilardi N and van Lookeren Campagne M: CRIg: A macrophage complement receptor required for phagocytosis of circulating pathogens. Cell. 124:915–927. 2006. View Article : Google Scholar : PubMed/NCBI

53 

Chow A, Schad S, Green MD, Hellmann MD, Allaj V, Ceglia N, Zago G, Shah NS, Sharma SK, Mattar M, et al: Tim-4+ cavity-resident macrophages impair anti-tumor CD8+ T cell immunity. Cancer Cell. 39:973–988.e979. 2021. View Article : Google Scholar : PubMed/NCBI

54 

Li J, Diao B, Guo S, Huang X, Yang C, Feng Z, Yan W, Ning Q, Zheng L, Chen Y and Wu Y: VSIG4 inhibits proinflammatory macrophage activation by reprogramming mitochondrial pyruvate metabolism. Nat Commun. 8:13222017. View Article : Google Scholar : PubMed/NCBI

55 

Vogt L, Schmitz N, Kurrer MO, Bauer M, Hinton HI, Behnke S, Gatto D, Sebbel P, Beerli RR, Sonderegger I, et al: VSIG4, a B7 family-related protein, is a negative regulator of T cell activation. J Clin Invest. 116:2817–2826. 2006. View Article : Google Scholar : PubMed/NCBI

56 

Srirussamee K, Mobini S, Cassidy NJ and Cartmell SH: Direct electrical stimulation enhances osteogenesis by inducing Bmp2 and Spp1 expressions from macrophages and preosteoblasts. Biotechnol Bioeng. 116:3421–3432. 2019. View Article : Google Scholar : PubMed/NCBI

57 

Wei J, Marisetty A, Schrand B, Gabrusiewicz K, Hashimoto Y, Ott M, Grami Z, Kong LY, Ling X, Caruso H, et al: Osteopontin mediates glioblastoma-associated macrophage infiltration and is a potential therapeutic target. J Clin Invest. 129:137–149. 2019. View Article : Google Scholar : PubMed/NCBI

58 

Su X, Xu BH, Zhou DL, Ye ZL, He HC, Yang XH, Zhang X, Liu Q, Ma JJ, Shao Q, et al: Polymorphisms in matricellular SPP1 and SPARC contribute to susceptibility to papillary thyroid cancer. Genomics. 112:4959–4967. 2020. View Article : Google Scholar : PubMed/NCBI

59 

Chen P, Zhao D, Li J, Liang X, Li J, Chang A, Henry VK, Lan Z, Spring DJ, Rao G, et al: Symbiotic macrophage-glioma cell interactions reveal synthetic lethality in PTEN-null glioma. Cancer Cell. 35:868–884.e866. 2019. View Article : Google Scholar : PubMed/NCBI

60 

Zhang Y, Du W, Chen Z and Xiang C: Upregulation of PD-L1 by SPP1 mediates macrophage polarization and facilitates immune escape in lung adenocarcinoma. Exp Cell Res. 359:449–457. 2017. View Article : Google Scholar : PubMed/NCBI

61 

Hurley PJ, Hughes RM, Simons BW, Huang J, Miller RM, Shinder B, Haffner MC, Esopi D, Kimura Y, Jabbari J, et al: Androgen-regulated SPARCL1 in the tumor microenvironment inhibits metastatic progression. Cancer Res. 75:4322–4334. 2015. View Article : Google Scholar : PubMed/NCBI

62 

Naschberger E, Liebl A, Schellerer VS, Schütz M, Britzen-Laurent N, Kölbel P, Schaal U, Haep L, Regensburger D, Wittmann T, et al: Matricellular protein SPARCL1 regulates tumor microenvironment-dependent endothelial cell heterogeneity in colorectal carcinoma. J Clin Invest. 126:4187–4204. 2016. View Article : Google Scholar : PubMed/NCBI

63 

Zhao SJ, Jiang YQ, Xu NW, Li Q, Zhang Q, Wang SY, Li J, Wang YH, Zhang YL, Jiang SH, et al: SPARCL1 suppresses osteosarcoma metastasis and recruits macrophages by activation of canonical WNT/β-catenin signaling through stabilization of the WNT-receptor complex. Oncogene. 37:1049–1061. 2018. View Article : Google Scholar : PubMed/NCBI

64 

Wu L, Sun S, Qu F, Sun M, Liu X, Sun Q, Cheng L, Zheng Y and Su G: CXCL9 influences the tumor immune microenvironment by stimulating JAK/STAT pathway in triple-negative breast cancer. Cancer Immunol Immunother. 72:1479–1492. 2023. View Article : Google Scholar : PubMed/NCBI

65 

Yang M, Lu J, Zhang G, Wang Y, He M, Xu Q, Xu C and Liu H: CXCL13 shapes immunoactive tumor microenvironment and enhances the efficacy of PD-1 checkpoint blockade in high-grade serous ovarian cancer. J Immunother Cancer. 9:e0011362021. View Article : Google Scholar : PubMed/NCBI

66 

Marcovecchio PM, Thomas G and Salek-Ardakani S: CXCL9-expressing tumor-associated macrophages: New players in the fight against cancer. J Immunother Cancer. 9:e0020452021. View Article : Google Scholar : PubMed/NCBI

67 

Xie Y, Chen Z, Zhong Q, Zheng Z, Chen Y, Shangguan W, Zhang Y, Yang J, Zhu D and Xie W: M2 macrophages secrete CXCL13 to promote renal cell carcinoma migration, invasion, and EMT. Cancer Cell Int. 21:6772021. View Article : Google Scholar : PubMed/NCBI

68 

Zhao S, Mi Y, Guan B, Zheng B, Wei P, Gu Y, Zhang Z, Cai S, Xu Y, Li X, et al: Tumor-derived exosomal miR-934 induces macrophage M2 polarization to promote liver metastasis of colorectal cancer. J Hematol Oncol. 13:1562020. View Article : Google Scholar : PubMed/NCBI

69 

Huang D, Sun W, Zhou Y, Li P, Chen F, Chen H, Xia D, Xu E, Lai M, Wu Y and Zhang H: Mutations of key driver genes in colorectal cancer progression and metastasis. Cancer Metastasis Rev. 37:173–187. 2018. View Article : Google Scholar : PubMed/NCBI

70 

Popat S, Hubner R and Houlston RS: Systematic review of microsatellite instability and colorectal cancer prognosis. J Clin Oncol. 23:609–618. 2005. View Article : Google Scholar : PubMed/NCBI

71 

Ward R, Meagher A, Tomlinson I, O'Connor T, Norrie M, Wu R and Hawkins N: Microsatellite instability and the clinicopathological features of sporadic colorectal cancer. Gut. 48:821–829. 2001. View Article : Google Scholar : PubMed/NCBI

72 

García-Marín R, Reda S, Riobello C, Cabal VN, Suárez-Fernández L, Vivanco B, Álvarez-Marcos C, López F, Llorente JL and Hermsen MA: Prognostic and therapeutic implications of immune classification by CD8(+) tumor-infiltrating lymphocytes and PD-L1 expression in sinonasal squamous cell carcinoma. Int J Mol Sci. 22:69262021. View Article : Google Scholar : PubMed/NCBI

73 

Zhang Y, Wang X, Shi M, Song Y, Yu J and Han S: Programmed death ligand 1 and tumor-infiltrating CD8+ T lymphocytes are associated with the clinical features in meningioma. BMC Cancer. 22:11712022. View Article : Google Scholar : PubMed/NCBI

74 

Chen YP, Zhang Y, Lv JW, Li YQ, Wang YQ, He QM, Yang XJ, Sun Y, Mao YP, Yun JP, et al: Genomic analysis of tumor microenvironment immune types across 14 solid cancer types: Immunotherapeutic implications. Theranostics. 7:3585–3594. 2017. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

December-2024
Volume 28 Issue 6

Print ISSN: 1792-1074
Online ISSN:1792-1082

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Feng Y, Cheng Z, Gao J, Huang T, Wang J, Tang Q, Pu K and Liu C: Revolutionizing prognostic predictions in colorectal cancer: Macrophage‑driven transcriptional insights from single‑cell RNA sequencing and gene co‑expression network analysis. Oncol Lett 28: 587, 2024.
APA
Feng, Y., Cheng, Z., Gao, J., Huang, T., Wang, J., Tang, Q. ... Liu, C. (2024). Revolutionizing prognostic predictions in colorectal cancer: Macrophage‑driven transcriptional insights from single‑cell RNA sequencing and gene co‑expression network analysis. Oncology Letters, 28, 587. https://doi.org/10.3892/ol.2024.14721
MLA
Feng, Y., Cheng, Z., Gao, J., Huang, T., Wang, J., Tang, Q., Pu, K., Liu, C."Revolutionizing prognostic predictions in colorectal cancer: Macrophage‑driven transcriptional insights from single‑cell RNA sequencing and gene co‑expression network analysis". Oncology Letters 28.6 (2024): 587.
Chicago
Feng, Y., Cheng, Z., Gao, J., Huang, T., Wang, J., Tang, Q., Pu, K., Liu, C."Revolutionizing prognostic predictions in colorectal cancer: Macrophage‑driven transcriptional insights from single‑cell RNA sequencing and gene co‑expression network analysis". Oncology Letters 28, no. 6 (2024): 587. https://doi.org/10.3892/ol.2024.14721