Skip to main content
Erschienen in: Journal of Translational Medicine 1/2022

Open Access 01.12.2022 | Research

Construction of a predictive model for immunotherapy efficacy in lung squamous cell carcinoma based on the degree of tumor-infiltrating immune cells and molecular typing

verfasst von: Lingge Yang, Shuli Wei, Jingnan Zhang, Qiongjie Hu, Wansong Hu, Mengqing Cao, Long Zhang, Yongfang Wang, Pingli Wang, Kai Wang

Erschienen in: Journal of Translational Medicine | Ausgabe 1/2022

insite
INHALT
download
DOWNLOAD
print
DRUCKEN
insite
SUCHEN

Abstract

Background

To construct a predictive model of immunotherapy efficacy for patients with lung squamous cell carcinoma (LUSC) based on the degree of tumor-infiltrating immune cells (TIIC) in the tumor microenvironment (TME).

Methods

The data of 501 patients with LUSC in the TCGA database were used as a training set, and grouped using non-negative matrix factorization (NMF) based on the degree of TIIC assessed by single-sample gene set enrichment analysis (GSEA). Two data sets (GSE126044 and GSE135222) were used as validation sets. Genes screened for modeling by least absolute shrinkage and selection operator (LASSO) regression and used to construct a model based on immunophenotyping score (IPTS). RNA extraction and qPCR were performed to validate the prognostic value of IPTS in our independent LUSC cohort. The receiver operating characteristic (ROC) curve was constructed to determine the predictive value of the immune efficacy. Kaplan–Meier survival curve analysis was performed to evaluate the prognostic predictive ability. Correlation analysis and enrichment analysis were used to explore the potential mechanism of IPTS molecular typing involved in predicting the immunotherapy efficacy for patients with LUSC.

Results

The training set was divided into a low immune cell infiltration type (C1) and a high immune cell infiltration type (C2) by NMF typing, and the IPTS molecular typing based on the 17-gene model could replace the results of the NMF typing. The area under the ROC curve (AUC) was 0.82. In both validation sets, the IPTS of patients who responded to immunotherapy were significantly higher than those who did not respond to immunotherapy (P = 0.0032 and P = 0.0451), whereas the AUC was 0.95 (95% CI = 1.00–0.84) and 0.77 (95% CI = 0.58–0.96), respectively. In our independent cohort, we validated its ability to predict the response to cancer immunotherapy, for the AUC was 0.88 (95% CI = 1.00–0.66). GSEA suggested that the high IPTS group was mainly involved in immune-related signaling pathways.

Conclusions

IPTS molecular typing based on the degree of TIIC in the TME could well predict the efficacy of immunotherapy in patients with LUSC with a certain prognostic value.
Hinweise

Supplementary Information

The online version contains supplementary material available at https://​doi.​org/​10.​1186/​s12967-022-03565-7.
Lingge Yang and Shuli Wei contributed equally to this work

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Abkürzungen
AUC
Area under the curve
CTLA-4
Cytotoxic T lymphocyte-associated antigen-4
CR
Complete response
DEGs
Differentially expressed genes
DFS
Disease-free survival
FDR
False discovery rate
GDSC
Genomics of drug sensitivity in cancer
GEO
Gene expression omnibus
GO
Gene ontology
GSEA
Gene set enrichment analysis
IC50
50% inhibitory concentration
ICIs
Immune checkpoint inhibitors
IPTS
Immunophenotyping score
irAEs
Immune-related adverse events
JAK/STAT
Janus kinase/signal transducer and activator of transcription
KEGG
Kyoto encyclopedia of genes and genomes
LASSO
Least absolute shrinkage and selection operator
LUSC
Lung squamous cell carcinoma
MAE
Mean absolute error
MAPK
Mitogen-activated protein kinase
MHC
Major histocompatibility complex
mTOR
Mammalian target of rapamycin
NES
Normalized enrichment score
NF-κB
Nuclear transcription factor-κB
NK
Natural killer
NMF
Non-negative matrix factorization
NP
Normalized P-value
NSCLC
Non-small cell lung cancer
OS
Overall survival
PCA
Principal component analysis
PD
Progression of disease
PD-1
Programmed cell death protein-1
PD-L1
Programmed cell death ligand-1
PERK
Protein kinase R-like endoplasmic reticulum kinase
PFS
Progression-free survival
PR
Partial response
RECIST
Response evaluation criteria in solid tumours
ROC
Receiver operating characteristic
SD
Stable disease
ssGSEA
Single-sample gene set enrichment analysis
TCGA
The Cancer Genome Atlas
TIDE
Tumor Immune Dysfunction and Exclusion
TIICs
Tumor-infiltrating immune cells
TIS
Tumor Inflammation Signature
TLR
Toll-like receptor
TMB
Tumor mutational burden
TNF
Tumor necrosis factor
TPM
Transcripts per kilobase million
tSNE
T-distributed stochastic neighbor embedding

Background

Lung cancer is a common malignant tumor worldwide. According to the 2020 global cancer statistics, the mortality and incidence rates of lung cancer rank first and second, respectively [1]. Lung squamous cell carcinoma (LUSC) is the second most common histological subtype of lung cancer with ~ 30% of all cases [2]. Due to the insidious onset and low early diagnosis rate, many patients with LUSC have already passed the opportunity for surgery by the time of diagnosis [3]. The 5-year survival rate of patients with LUSC who receive surgery is still low at 12.4% [4]. Compared with lung adenocarcinoma, LUSC has a low rearrangement rate of EGFR gene mutation and ALK fusion gene, and strong tumor heterogeneity [5], Therefore, LUSC is limited in gene mutation-based targeted therapy applications [6, 7]. Other treatments such as chemotherapy and radiotherapy also have a limited impact on the long-term survival of patients with LUSC [8]. Thus, patients with LUSC generally have a poor prognosis [9].
In clinical application, immunotherapy plays an integral anti-tumor role by activating the immune system and is rapidly becoming an important tool for cancer treatment. The most widely used immunotherapy is immune checkpoint inhibitors (ICIs), and they have shown promising therapeutic outcomes in non-small cell lung cancer (NSCLC) [10]. However, the response rate of immunotherapy is relatively low, and only a subset of patients show meaningful clinical response or benefit [11]. As a target of PD-1/PD-L1 antibodies, the PD-L1 level in cancer cells as measured by immunohistochemistry is the only FDA-approved and widely used biomarker for predicting response to ICIs in clinical practice. However, the predictive ability of the PD-L1 level is limited, and despite a high PD-L1 level, a proportion of patients receiving ICIs still do not respond; similarly, a negative PD-L1 level also does not reliably preclude a response to PD-1/PD-L1 blockade [12], suggesting there is an urgent need for effective biomarkers capable of screening patients with LUSC according to their likelihood of benefiting from ICI therapy. Beyond the intrinsic factors of tumor cells, studies have identified the tumor microenvironment (TME) characteristics also determine the ICI tumor response [13]. Among them, immune cells play key roles in mediating immune surveillance and regulating tumor growth [14]. Therefore, tumor-infiltrating immune cells (TIICs) may be a potential biomarker to predict the efficacy of immunotherapy.
A clinical prediction model is a tool that combines multiple predictors to evaluate the probability of an individual presenting with a certain disease or clinical outcome. Some clinical prediction models have potential value for screening, diagnosis, treatment, and prognostic prediction of lung cancer [1517]. With the rapid development of high-throughput sequencing and bioinformatics analysis methods, obtaining cancer-related genomes, transcriptomes, and immune-related information has become readily easier. This has enabled the construction of lung cancer prediction models based on gene-related predictors, which are now widely used in clinical practice.
At present, there is a relative lack of predictive models for the efficacy of immunotherapy in LUSC based on TIIC. Our study intends to construct a predictive model for the efficacy of immunotherapy for patients with LUSC based on the degree of TIIC. First, non-negative matrix factorization (NMF) [18] was used to classify the gene expression profile of patients with LUSC from The Cancer Genome Atlas (TCGA) database. Then, after intersecting differentially expressed genes (DEGs) between NMF typing, survival-related genes, and their comparison with two validation gene sets of patients receiving immunotherapy, a least absolute shrinkage and selection operator (LASSO) analysis was performed [19]. Finally, 17 genes were screened out and the corresponding regression coefficients were obtained, which were used to construct an immunophenotyping score (IPTS) molecular typing, and used to analyze the predictive value of IPTS on the efficacy of immunotherapy for patients with LUSC.

Method

Data collection and processing

The clinical information and gene expression profile matrix of patients with LUSC were downloaded from the TCGA database (https://​cancergenome.​nih.​gov, access date: October 15, 2021). A total of 501 samples with complete clinical information and expression profile matrix were selected as the training set to construct the immune efficacy prediction model. Then, the gff3 file (v37, released on October 14, 2021) was downloaded from GENCODE (https://​www.​gencodegenes.​org/​human/​) [20], the Gene Symbol and ENSG_ID extracted using R v4.1.2, and matched with the TCGA-LUSC expression profile matrix to convert ENSG_ID to Gene Symbol. Next, the count data were transformed into transcripts per kilobase million (TPM) data based on gene length for subsequent analyses.
In addition, the clinical information and gene expression matrix of two data sets of NSCLC with immunotherapeutic efficacy, GSE126044 [21] and GSE135222 [22, 23], were downloaded from the gene expression omnibus (GEO) (https://​www.​ncbi.​nlm.​nih.​gov/​geo, access date: February 11, 2022) database as the validation sets. The GSE126044 dataset was sequenced using the HiSeq 2500 (GPL16791; Illumina, San Diego, CA, USA) platform, with a total of 16 NSCLC samples. The GSE135222 dataset was also sequenced using the GPL16791 platform, with a total of 27 samples, and this dataset contained the prognosis information of progression-free survival (PFS).

Single-sample gene set enrichment analysis (ssGSEA) to assess the degree of immune cells infiltration

According to the gene signatures of 28 types of immune cells reported by Jia Q [24], we used the “GSVA” package (v1.42.0) [25] and the ssGSEA method to obtain the enrichment scores of 28 types of immune cells in each of the 501 LUSC cases in the training set. These 28 kinds of immune cells can be divided into cell types executing anti-tumor immunity (including activated CD4 T, activated CD8 T, activated dendritic, CD56 bright natural killer (NK), central memory CD4 T, central memory CD8 T, effector memory CD4 T, effector memory CD8 T, NK, NK T, type 1 T helper, and type 17 T helper cells), cell types executing pro-tumor, immune suppressive function (including CD56 dim NK, immature dendritic, myeloid-derived suppressor, plasmacytoid dendritic, regulatory T, type 2 T helper cells, neutrophils, and macrophages), and other cell types (activated B, gamma delta T, immature B, mast, memory B, T follicular helper cells, eosinophils, and monocytes).

NMF typing

After normalizing the above matrix of immune cell enrichment scores, we used the “NMF” package (v0.23.0) [26] for typing with rank set to 2:10, the method to brunet, and nrun to 100. Then, we used the “Rtsne” package (v0.15) [27] and the “prcomp” function of the “stats” package (v3.6.0) [28] to perform dimensionality reduction analysis to verify the feasibility of the NMF typing results. In addition, to clarify the differences in TIIC between different types, we performed correlation analysis and difference analysis on the anti-tumor immunity enrichment scores and the pro-tumor immunity enrichment score between different subtypes, and analyzed the difference in the enrichment scores of 28 kinds of TIIC respectively, to further demonstrate the reliability of NFM typing results.

DEGs of NMF typing and their functional enrichment analysis

To clarify the DEGs of different NMF types, we used the “limma” package (v3.50.1) [29] to analyze the DEG profile, and used the “p.adjust” function to calculate the significant false discovery rate (FDR, q-value) of each gene. FDR (q-value) < 0.05 was considered to be statistically significant. Then, the “clusterProfiler” package (v3.14.3) [30] was used for functional enrichment analysis of gene ontology (GO) and Kyoto encyclopedia of genes and genomes (KEGG) to obtain the results of gene set enrichment. The minimum gene set was 5 and the maximum gene set was 5,000, with P < 0.05 and FDR < 0.05 considered meaningful. The results were ranked by FDR and the top ten functional enrichment results were plotted.

Survival analysis and screening for genes affecting overall survival (OS) and disease-free survival (DFS)

To determine whether there is a difference in survival between patients with different molecular types, we grouped the patients according to the NMF types, and then used the “survival” package (v3.3-1) [31] for survival analysis, with the optimal cutoff value calculated using the “surv_cutpoint” function. Then, the “survminer” package (v0.4.9) was used for plotting survival curves [32]. In addition, to screen the genes that have an impact on the prognosis and survival of LUSC in the TCGA database, we first removed the samples with a survival time of fewer than 30 days, and then performed survival analyses on all genes in the gene expression profile to obtain the genes affecting OS and DFS, respectively.

Construction of an immune efficacy prediction model

Due to the different sequencing platforms or sequencing depths, many genes detected in the training set were not detected in the validation sets. To better use the validation set for verification, we used DEGs, genes affecting the OS and DFS of patients, and genes measured in the two validation sets of GSE126044 and GSE135222. After intersecting these five gene sets, the screened genes were used to construct the immune efficacy prediction model. A Venn diagram was plotted using the online tool Bioinformatics & Evolutionary Genomics (http://​bioinformatics.​psb.​ugent.​be/​webtools/​Venn).
The gene expression profile matrix screened above for modeling and NMF typing were extracted and then the LASSO regression model was constructed using the “glmnet” package (v4.1–3) [33, 34], with nfold set to 10 and λ equaling lambda.min. After obtaining the regression coefficients of the screened genes, the IPTS equation was constructed based on these coefficients. Then, nomograms and calibration curves were plotted using the “rms” package (v6.2-0) [35] to visualize the regression analysis results. In addition, we plotted a Sankey diagram using the “networkD3” package (v0.4) [36] to visualize the typing results and their corresponding gene signatures.

Our independent LUSC validation set collection and follow-up

All paraffin sections from 10 cases of LUSC tissues were collected in The Second Affiliated Hospital of Zhejiang University School of Medicine from November 2019 to February 2022. Clinicopathological characteristics and prognostic survival information of these LUSC patients, including ages, gender, TNM stage, clinical stage, tumor size before and after treatment, tumor site, best response evaluation and PD-L1 immunohistochemistry data were acquired. The follow-up date was ended at July 7, 2022, and outpatient and telephone follow-up were performed. This study was approved by the institutional review committee of The Second Affiliated Hospital of Zhejiang University School of Medicine (Approval Number: 2022-0548/I2022685). All the patients have written informed consent before surgery.

RNA extraction and real-time quantitative PCR (qPCR)

For real-time qPCR analysis, the BIOG RNA FFPE Tissue Kit (Baidai, Changzhou, China) was used to extract total RNA from 10 samples of Paraffin section of lung from patients receiving immunotherapy according to the manufacturer’s instructions. cDNA was synthesized using the HiScript III All-in-one RT SuperMix Perfect (Vazyme, Nanjing, China). Real-time q-PCR was performed to detect the expression of the screened genes using TB Green Premix Ex Taq II (Takara, Dalian, China) to calculate IPTS. Gene expression levels were normalized to the “housekeeping” gene GAPDH. The primers and their sequences were listed in Additional file 2: Table S1.

Reliability and verification of model for immunotherapy efficacy prediction

To verify the discrimination of NMF typing by the prediction model and whether it can replace the sample typing results, we first calculated the IPTS value of each sample in the training and two validation sets as well as our independent cohort according to both the gene expression value and the constructed IPTS equation. Then, the IPTS and NMF typing results of the training set were integrated, and the receiver operating characteristic (ROC) curves were constructed through the “pROC” package (v1.18.0) [37] to evaluate the ability of the prediction model to judge the NMF typing. Next, according to the cutoff value of the ROC curve, the training set was divided into high and low score groups. The differences and correlations between the high and low IPTSes for different NMF typing, as well as the pro- and anti-tumor immunity enrichment scores between the high and low score groups, were analyzed, respectively. In addition, the differences between the enrichment scores of 28 types of immune cells were also analyzed to clarify the degree of coincidence between the IPTS molecular typing results and the NMF typing results, i.e., whether the substitution of the IPTS molecular typing results for the NMF typing results is reasonable. Besides, since the genes in the model have an impact on the OS and DFS of the training set, we also performed survival analysis and plotted the survival curve to evaluate the prognostic predictive value of the model.
According to the clinical information of patients with LUSC in the TCGA database, almost no patients received immunotherapy. To evaluate whether the constructed immune typing model can predict immunotherapy efficacy, we validated the immunotherapy efficacy in the two validation sets for patients with NSCLC as well as our independent LUSC cohort. First, we compared whether there were significant differences in the IPTSes between groups of patients with lung cancer and different immune responses. Due to the small sample size in the GSE126044 dataset and the IPTS in this dataset with non-normally distributed and uneven variance, the Mann–Whitney rank-sum test was used for the differences between groups in this dataset, and P < 0.05 was considered to be statistically significant. And due to the power distribution of IPTSes in our independent LUSC cohort, log2 transformation was conducted before using the student’s t-test. Then, the ROC curve was constructed according to the IPTS and immune response results to evaluate the predictive value of the immune efficacy of the model. In addition, since the GSE135222 dataset and our cohort contain data on the PFS of patients, we further conducted survival analyses to verify the predictive value of immune efficacy and evaluate the prognostic prediction ability of this model.

Efficacy prediction of other anti-tumor drugs

Genomics of drug sensitivity in cancer (GDSC; https://​www.​cancerrxgene.​org, access date: February 27, 2022) [38] contains the sequencing data of more than 1000 human tumor cell lines and the treatment results of tumor cells by more than 100 anti-tumor drugs, which facilitate finding molecular characteristics of tumors and predicting the response of targets to anti-tumor drugs. The sequencing results of all cell lines in the database and the 50% inhibitory concentration (IC50) of cell lines treated with anti-tumor drugs were downloaded, and the results of all 15 cell lines of LUSC sequenced in this database were extracted. Next, the IPTS of the 15 cell lines were calculated and then divided into two groups: 8 cases with high IPTS and 7 cases with low IPTS. The differences in IC50 of anti-tumor drugs between the two groups were tested, and the anti-tumor drugs with statistical significance (P < 0.05) were selected.

Analysis of differences and correlation between two IPTS groups in immune microenvironment score and immune molecular typing

Through the “estimate” package (v1.0.13) [39], we evaluated the three immune microenvironment related scores of 501 samples in the training set as well as analyzed the differences and correlations between the high and low IPTS groups. Meanwhile, according to the summary of genotype and immunophenotype by Charoentong [40] and Hu [41], we obtained the following five genetic markers of immune molecular typing, namely chemokines, receptors, major histocompatibility complex (MHC) molecules, immuno-inhibitors, and immuno-stimulators. Next, the enrichment scores of the above five immune molecular typing in the training set were calculated by ssGSEA, and the differences between the high and low score groups were then analyzed. As the current clinically used ICIs are mainly anti-CTLA-4 and anti-PD-1/PD-L1 antibodies, we analyzed the differences and correlations between the expression of four immune checkpoints CTLA-4, PD-1 (PDCD1), and its two ligands PD-L1 (CD274) and PDL-2 (PDCD1LG2) in the training set between high and low score groups.

Gene set enrichment analysis (GSEA)

We performed GSEA (https://​www.​gsea-msigdb.​org/​gsea, access date: March 1, 2022) [42] on the gene expression profile of the training set based on the high and low IPTS groups. First, the subset of c2.cp.kegg.v7.4.symbols.gmt were downloaded to assess relevant pathways and molecular mechanisms. Based on the gene expression profile and IPTS grouping, the minimum gene set was 5, the maximum gene set was 5000, and the number of resampling was set to 1000. Then, ranked by the normalized enrichment score (NES), the top seven results were visualized. A normalized P-value (NP) < 0.05 was considered to be meaningful.

Statistical analysis

All data in this study were analyzed and plotted using R v4.1.2 and Prism v8.0.1 (GraphPad, San Diego, CA, USA). The parameters not mentioned in the methods were all default parameters, and the data visualization not mentioned was all plotted by the “ggplot2” package (v3.3.5) [43]. Continuous variables were displayed as mean ± standard deviation. Unless mentioned otherwise mentioned, the student’s t-test was used to compare the differences between the two groups. The differences between groups of discrete variables were analyzed using the chi-squared test. The Pearson test was used for correlation analysis, and the log-rank test was used for survival analysis. P < 0.05 was considered as a statistically significant difference (*P < 0.05, **P < 0.01, ***P < 0.001, ****P < 0.0001).

Results

NMF typing divides the training set into low and high immune cell infiltration types

The clinical information of the training, two validation sets and our independent LUSC cohort is detailed in Additional file 3: Table S2. NMF was used to classify 501 patients with LUSC in the training set. It was found that when the rank value was 2–3, the cophenetic typing index decreased the most (Fig. 1A). Therefore, a rank value of 2 was selected, and the patients were divided into a low immune cell infiltration type (cluster 1; C1) and a high immune cell infiltration type (cluster 2; C2). The typing efficacy of other NMF indicators is shown in Additional file 1: Figure S1. The heat map also showed that when the number of types was limited to two, the samples of the training set could be well distinguished (Fig. 1B). The discriminatory capacity when more subtypes were used in the training can be seen in Additional file 1: Fig. S2. In addition, dimensionality reduction by t-distributed stochastic neighbor embedding (tSNE) (Fig. 1C) and principal component analysis (PCA) (Fig. 1D) showed that C1 and C2 had good discrimination ability, suggesting the feasibility of using this classification.
There were significant positive correlations between the scores of cell types executing anti-tumor immunity and those executing pro-tumor immunity in both the C1 and C2 patients (Fig. 1E), with a correlation coefficient of 0.86 in C1 patients (P = 1.2e−56) and 0.83 in C2 patients (P = 2.8e−82). The enrichment score of the anti- (P < 0.0001) and pro-tumor immunity (P < 0.0001) were significantly higher in C2 patients than in C1 patients (Fig. 1F). In addition, there were significant differences in the enrichment scores of 28 types of immune cells, with significantly lower scores in type C1 than type C2 (Fig. 1G). It is suggested that type C1 of LUSC can be regarded as a “cold tumor” with low infiltration of immune cells, while type C2 tends to be a “hot tumor”, i.e., a tumor with high infiltration of immune cells. These results reveal that NMF typing had a strong ability to distinguish the degree of immune cell infiltration in the training set.

Clinicopathological characteristics between NMF types

According to the NMF typing, we performed difference analyses on several important clinicopathological characteristics of the patients in the training set, including age, gender, TNM stage, and clinical stage. The results only showed significant differences in gender between the two subtypes. A total of 36 patients in type C1 were female (7.19%), which was significantly less than the 94 patients (18.76%) in type C2 (P = 0.0112). There were no differences in other clinicopathological characteristics such as age and stage between the subtypes (Table 1). Subsequent survival analysis revealed no significant difference between the two subtypes for either the OS (P = 0.74, Fig. 2A) or DFS (P = 0.5, Fig. 2B) for patients with LUSC.
Table 1
Correlation between clinicopathological characteristics and NMF typing in the TCGA database
Characteristics
C1 (N = 185)
C2 (N = 316)
Total (N = 501)
P-value
Age
 Mean ± standard deviation
66.33 ± 8.74
67.71 ± 8.46
67.20 ± 8.58
0.0836
 Median [min, max]
68 [39, 85]
68 [40, 90]
68 [39, 90]
 
Gender
 Female
36 (7.19%)
94 (18.76%)
130 (25.95%)
0.0112*
 Male
149 (29.74%)
222 (44.31%)
371 (74.05%)
T
 T1
36 (7.19%)
78 (15.57%)
114 (22.75%)
0.5468
 T2
112 (22.36%)
181 (36.13%)
293 (58.48%)
 T3
29 (5.79%)
42 (8.38%)
71 (14.17%)
 T4
8 (1.60%)
15 (2.99%)
23 (4.59%)
N
 N0
112 (22.36%)
207 (41.32%)
319 (63.67%)
0.1161
 N1
57 (11.38%)
74 (14.77%)
131 (26.15%)
 N2
16 (3.19%)
24 (4.79%)
40 (7.98%)
 N3
0
5 (1.00%)
5 (1.00%)
 Nx
0
6 (1.20%)
6 (1.20%)
 
M
 M0
149 (29.74%)
262 (52.30%)
411 (82.04%)
0.4213
 M1
1 (0.20%)
6 (1.20%)
7 (1.40%)
 Mx
34 (6.79%)
45 (8.97%)
79 (15.76%)
 
 Unknown
1 (0.20%)
3 (0.60%)
4 (0.80%)
 
Stage
 Stage I
80 (15.97%)
164 (32.73%)
244 (48.70%)
0.1160
 Stage II
68 (13.57%)
94 (18.76%)
162 (32.34%)
 Stage III
35 (6.99%)
49 (9.78%)
84 (16.77%)
 Stage IV
1 (0.20%)
6 (1.20%)
7 (1.40%)
 Unknown
1 (0.20%)
3 (0.60%)
4 (0.80%)
 
IPTS groups
 High score
52 (10.38%)
254 (50.70%)
306 (61.08%)
< 0.0001****
 Low score
133 (26.55%)
62 (12.38%)
195 (38.92%)
Bold values indicate P < 0.05 and *P < 0.05, ****P < 0.0001

DEGs between NMF types are mainly involved in immune system regulation

Through differential analysis, 468 genes were significantly up-regulated in type C1 compared with type C2, and 2179 genes were significantly down-regulated (Fig. 2C). GO enrichment analysis showed that DEGs were mainly enriched in the following biological processes: immune system process (48.50%), immune response (42.02%), regulation of immune system process (27.84%), etc. (Fig. 2D); In terms of cellular components, they were mainly enriched in the extracellular region (37.87%), intrinsic components of the membrane (15.40%), and integral components of the membrane (14.78%), etc. (Fig. 2E); while in terms of molecular functions, these genes were mainly enriched in signaling receptor activity (16.61%), molecular transducer activity (16.61%), and signaling receptor binding (15.25%), etc. (Fig. 2F). KEGG enrichment analysis showed that DEGs were mainly involved in cytokine-cytokine receptor interaction (12.75%), chemokine signaling pathway (8.55%), viral proteins, viral protein interaction with cytokine and cytokine receptor (7.46%), and other pathways (Fig. 2G). Additional file 4: Table S3 shows all meaningful GO and KEGG enrichment analysis results.

Construction of an immune infiltration prediction model based on 17 genes

Although NMF typing could better distinguish the abundance of TIICs, it could not predict the survival prognosis of patients. Therefore, a new predictive model needed to be developed on this premise. A total of 20 genes were obtained for constructing the prediction model, by taking the intersection of the following five gene sets: DEGs, genes affecting the OS and DFS of patients, and genes sequenced in the two validation sets (Fig. 3A). Through LASSO regression, a total of 17 genes with a regression coefficient were selected (Fig. 3B). In addition, it can be seen from Fig. 3C that the model had a higher ROC area under the curve (AUC) value when considering the minimum value of the tuning parameter (λ).
Based on these 17 genes and their regression coefficients, the IPTS model was constructed as follows: IPTS = 0.4869250211 − 0.1428834537 × AKAP2 expression value − 0.12060842 × NANOG expression value – 0.0951070744 × TMEM236 expression value − 0.0436119966 × NTSR1 expression value − 0.0258542814 × LRRC38 expression value − 0.0170225681 × GCGR expression value − 0.0011330363 × MARCO expression value − 0.0008511336 × PF4 expression value − 0.0004418332 × RP1 expression value + 0.0023249088 × ALOX5 expression value + 0.0021763779 × FCGR2A expression value + 0.0006362408 × KCNQ3 expression value + 0.0247048306 × NLRP12 expression value + 0.0314720069 × SCARF1 expression value + 0.0013954206 × SIGLEC12 expression value + 0.0004957628 × TGM2 expression value + 0.0617891897 × VSTM1 expression value.
In addition, the predictive effect of the prognostic model on type C2 was visualized by constructing a nomogram (Fig. 3D). It can be seen that when IPTS = 0.6369, the probability of patients in type C1 and C2 was 50%, and the probability of patients in type C2 was higher when the value exceeded 0.6369. Besides, under the condition of 1000 repetitions, the mean absolute error (MAE) of the calibration curve (Fig. 3E) was 0.062, and the curve fitting was suitable, indicating a sound prediction effect. As can be seen from this model, there were nine genes with regression coefficient < 0—which can be used as gene signatures of type C1; and eight genes with regression coefficient > 0—which can be used as gene signatures of type C2 (Fig. 3F). Thus, molecular typing was achieved to a certain extent.

Molecular typing based on IPTS prediction model can replace NMF typing

Through analysis and drawing the ROC curve, the AUC was found to be 0.82 (95% CI = 0.86–0.79) (Fig. 4A), and the cutoff value corresponding to the maximum Youden index showed a prediction probability of 50% in the nomogram, i.e., 0.6369. At this point, the prediction sensitivity of the ROC curve was 0.8038, and the specificity was 0.7189, indicating that the IPTS model could well predict the NMF typing of patients with LUSC. Moreover, the IPTS of patients in type C2 was significantly higher than that of patients in type C1 (P = 0.0026, Fig. 4B), and the number of patients in type C2 with higher IPTS was significantly higher than that of patients in type C1 (P < 0.0001, Table 1). It can be preliminarily concluded that molecular typing based on the IPTS could well predict NMF typing.
To further clarify whether IPTS subtyping could replace NMF typing, we also performed difference analysis and correlation analysis. As shown in Fig. 4C, the enrichment scores of anti-tumor immunity (P < 0.0001) and pro-tumor immunity (P < 0.0001) in patients with high IPTS were significantly higher than those in patients with low IPTS. Furthermore, there were significant correlations between the enrichment scores of anti- and pro-tumor immunity both in the high and low IPTS groups (Fig. 4D). The correlation coefficient of the high score group was 0.86 (P = 5.1e−89), while the correlation coefficient of the low score group was 0.90 (P = 1.8e−73). In addition, except for CD56 bright NK cells (P = 0.07), the enrichment scores of TIICs in the high IPTS group were significantly higher than those in the low IPTS groups (P < 0.0001, Fig. 4E). Therefore, molecular typing based on IPTS can replace NMF typing and has potential therapeutic value for clinical application.
Since our prediction model was constructed based on genes that affect the OS and DFS of patients, we performed survival analysis to evaluate the prognostic predictive value of this model. Furthermore, we assessed whether the model could compensate for the missing prognostic prediction function of the NMF typing. The results suggested that when the optimal cutoff value was used instead of the cutoff value of the ROC curve, the low IPTS group had better OS (HR = 1.06, 95% CI = 0.98–1.16, P = 0.03, Fig. 4F) and DFS (HR = 1.12, 95% CI = 1.03–1.20, P = 0.0027, Fig. 4G) than the high IPTS group. The prediction model could predict the prognosis of patients under certain conditions. For example, when predicting OS, the best cutoff value of IPTS was 0.75526; while when predicting DFS, the best cutoff value of IPTS was 0.96915 in the training set.

Immune infiltration prediction model predicts immune efficacy of immunotherapy and the potential therapeutic value of five anti-tumor drugs

The above survival analysis results indicated that patients with high IPTS have worse prognoses. However, to our knowledge, these patients theoretically benefit from immunotherapy due to the high degree of immune cell infiltration. Therefore, we analyzed and verified this view via two data sets in patients that received immunotherapy. In the NSCLC cohort (GSE126044) receiving anti-PD-1 antibody immunotherapy, patients who responded to immunotherapy had significantly higher IPTSes than patients who did not respond to immunotherapy (P = 0.0032) (Fig. 5A) with a ROC AUC of 0.95 (95% CI = 1.00–0.84), while the ROC AUC of PD-L1 was 0.73 (95% CI = 0.99–0.46, Fig. 5D) indicating that IPTS has a great predictive effect for the efficacy of immunotherapy in this dataset. In another NSCLC cohort (GSE135222), patients who benefited from immunotherapy had higher IPTSes than those who did not benefit from immunotherapy (P = 0.0451) (Fig. 5B) with a ROC AUC of 0.77 (95% CI = 0.96–0.58), and it was also larger than that of PD-L1 of 0.69 (95% CI = 0.90–0.48, Fig. 5E), suggesting that the IPTS has a good predictive value of immunotherapy efficacy in this cohort. In addition, this dataset reported the prognostic information on the PFS of patients. By taking the optimal cutoff value (IPTS = − 2.13), the PFS of patients with high IPTS after immunotherapy was better than that of patients with low IPTS (HR = 0.72, 95% CI = 0.5–1.04, P = 0.0059, Fig. 5G), which also showed a good predictive value of immunotherapy efficacy and a certain predictive value of survival prognosis.
In our independent LUSC cohort, there are two patients received complete response (CR), 3 patients received partial response (PR), 4 patients received stable disease (SD), and 1 patient received progression of disease (PD) according to response evaluation criteria in solid tumours (RECIST, v1.1). All ten patients were divided into two groups, responder group (CR + PR) and non-responder group (SD + PD) (Additional file 1: Fig. S3). The results showed that patients in responder group had significantly higher IPTSes than patients in non-responder group (P = 0.0325) (Fig. 5C) with a ROC AUC of 0.88 (95% CI = 1.00–0.66), which was larger than that of PD-L1 of 0.64 (95% CI = 1.00–0.26, Fig. 5F). The PFS of patients with high IPTS after immunotherapy was better than that of patients with low IPTS (P = 0.0403, Fig. 5H) by taking the best cutoff value (IPTS = − 4.05). Therefore, it could be confirmed that the constructed immune infiltration prediction model has predictive value for immune efficacy, i.e., immunotherapeutic efficacy could be better for patients with high IPTS.
From the above analysis, patients with low IPTS were not likely to benefit from immunotherapy. For this subtype of patients, we initially screened other anti-tumor drugs that may have a curative effect through the GDSC database. The analysis results showed that the IC50 of the five anti-tumor drugs acetalax (P = 0.0168), AZD2014 (P = 0.0416), GSK2606414 (P = 0.0145), obatoclax mesylate (P = 0.0061), and VSP34_8731 (P = 0.0163) were higher in the high IPTS group than in the low IPTS group in LUSC cell lines (Fig. 5I). Moreover, it can be seen from the heat map (Fig. 5J) that the IC50 value of the high IPTS group was generally higher than that of the low IPTS group, indicating that patients in the low IPTS group might be more sensitive to these five drugs.

IPTS positively correlates with immune microenvironment score and expression of immune-related genes signatures

The immune microenvironment scores of 501 patients with LUSC in the training set were evaluated using grouping analysis of IPTS molecular typing. The results showed that the stromal score (P < 0.0001), immune score (P < 0.0001), and ESTIMATE score (P < 0.0001) in the high IPTS group were significantly higher than those in the low IPTS group (Fig. 6A), suggesting that higher stromal cell levels and infiltration levels of immune cells in the high compared to low IPTS groups. This further confirms that tumors in the high IPTS subtype tended to be “hot tumors”. Moreover, through correlation analysis, IPTS was significantly positively correlated with the stromal score (P = 2.0e−22, r = 0.42, Additional file 1: Fig. S4A), immune score (P = 2.2e−31, r = 0.49, Additional file 1: Fig. S4B), and ESTIMATE score (P = 1.9e−30, r = 0.48, Additional file 1: Fig. S4C). Meanwhile, the correlation analysis of IPTS molecular typing showed that only the immune score (low score: P = 0.04, r = 0.14; High score: P = 8.4e−16, r = 0.44, Fig. 6E) was statistically significant in the low IPTS group, whereas there was no significant difference in stromal score (Low score: P = 0.21, r = 0.09; High score: P = 4.5e−9, r = 0.33, Fig. 6D) and ESTIMATE score (low score: P = 0.08, r = 0.13; High score: P = 1.2e−14, r = 0.42, Fig. 6F), suggesting that the three immune microenvironment scores were mainly positively correlated with IPTS in the high score group.
In addition, we analyzed the five immune molecular typing scores of the samples in the training set by ssGSEA. The detailed gene signatures of the immune molecular typing gene markers are listed in Additional file 5: Table S4. Similarly, according to the IPTS molecular typing analysis, the results showed that the enrichment scores of chemokines (P < 0.0001), receptors (P < 0.0001), MHC molecules (P < 0.0001), immuno-inhibitors (P < 0.0001), and immuno-stimulators (P < 0.0001) in the high IPTS group were significantly higher than those in the low IPTS group (Fig. 6B). Furthermore, the four targets of immunotherapy drugs commonly used in clinical practice all relate to immuno-inhibitors. Therefore, the differences in expression values of the four immuno-inhibitors CTLA-4, PD-1 (PDCD1), and its two ligands PD-L1 (CD274) and PDL-2 (PDCD1LG2) were analyzed between the IPTS groups in the training set. The results showed that the expressions of CTLA-4 (P < 0.0001), PD-1 (P < 0.0001), and PD-L2 (P < 0.0001) in the high IPTS group were significantly higher than those in the low IPTS group, whereas that of PD-L1 expression was insignificant (P = 0.22) (Fig. 6C). Further correlation analysis showed that IPTS was significantly positively correlated with the expressions of CTLA4 (P = 2.6e−11, r = 0.29, Additional file 1: Fig. S4D), PD-1 (P = 1.8e−8, r = 0.25, Additional file 1: Fig. S4E), PDL1 (P = 8.3e−3, r = 0.12, Additional file 1: Fig. S4F), and PD-L2 (P = 6.4e−9, r = 0.26, Additional file 1: Fig. S4G). The results of IPTS molecular typing correlation analysis showed a significant negative correlation of PD-1 expression with IPTS in the low score group (P = 0.04, r = − 0.15, Fig. 6H), whereas those of the other three immuno-inhibitors were not statistically significant. In the high score group, the expressions of the four immuno-inhibitors CTLA-4 (P = 2.7e−7, r = 0.29, Fig. 6G), PD-1 (P = 1.7e−5, r = 0.24, Fig. 6H), PD-L1 (P = 7.2e−4, r = 0.19, Fig. 6I), and PD-L2 (P = 3.4e−9, r = 0.33, Fig. 6J) were positively correlated with IPTS. To a certain extent, the above results further provide a theoretical basis for better immunotherapy efficacy in patients with high IPTS.

High IPTS involved in immune-related signaling pathways

The results of GSEA indicated that the IPTS was mainly related to Parkinson’s disease (NES = − 1.9186, NP = 0.0097), oxidative phosphorylation (NES = − 1.8862, NP = 0.0094), Huntington’s disease (NES = − 1.8586, NP = 0.0119), Alzheimer’s disease (NES = − 1.8144, NP = 0.0190), spliceosome (NES = − 1.7813, NP = 0.0215), homologous recombination (NES = − 1.7466, NP = 0.0214), nucleotide excision repair (NES = − 1.6533, NP = 0.0427), etc. (Fig. 7A). In contrast, high IPTS was mainly associated with cytokine-cytokine receptor interaction (NES = 2.6726, NP < 0.0001), chemokine signaling pathway (NES = 2.5505, NP < 0.0001), natural killer cell mediated cytotoxicity (NES = 2.5238, NP < 0.0001), leukocyte trans-endothelial migration (NES = 2.4652, NP < 0.0001), Leishmania infection (NES = 2.4424, NP < 0.0001), cell adhesion molecules (NES = 2.4418, NP < 0.0001), JAK-STAT signaling pathway (NES = 2.4409, NP < 0.0001), etc. (Fig. 7B). These results indicate that low IPTS was mainly involved in disease, genetic, and metabolic-related signaling pathways, while high IPTS was mainly involved in immune-related signal pathways. All the results of GSEA are detailed in Additional file 6: Table S5.

Discussion

The past decade witnessed great strides in cancer diagnosis and treatment. However, progress in improving the survival of patients with lung cancer has been slow, with an average 5-year survival rate of only 10–20% in most countries [44, 45]. In recent years, immunotherapy has achieved promising results in clinical practice. The latest research suggested a 5-year survival rate as high as 23.2% in patients with advanced NSCLC using anti-PD-1 antibodies as the first-line treatment [46]. Furthermore, the 5-year survival rate of patients treated with anti-PD-1 antibodies as a second-line treatment has also reached 16% [47], which is twice as high as that of traditional treatments. Nevertheless, several studies have revealed that only ~ 20% of patients with NSCLC could benefit from ICI therapy [48, 49], which illustrates the importance of selecting patients that will potentially benefit. Recently, Tian et al. [50] conducted an immune subgroup analysis study on NSCLC including LUSC, lung adenocarcinoma, and lung adenosquamous carcinoma, and found that mast cell types had a significant impact on the prognosis of patients with LUAD while the presence of monocytes was significantly associated with OS in patients with LUSC. Furthermore, the authors pointed out that LUSC and LUAD may require independent analysis. This is in accordance with a study reported by Jiang et al. [115] on the prediction of immunotherapy efficacy in NSCLC that also suggested the underlying immune response mechanism between LUAD and LUSC may be different. Therefore, we constructed a prediction model of immunotherapy efficacy to improve the accuracy of screening patients with LUSC for potential benefit from ICI treatment.
Detecting the expression level of PD-L1 is the most commonly used method to predict the efficacy of immunotherapy [51]. Some scholars have previously constructed some efficacy prediction models for tumor immunotherapy, such as the Tumor Immune Dysfunction and Exclusion (TIDE) [52] and the Tumor Inflammation Signature (TIS) [53, 54]. By comparing with PD-L1 expression level to predict the efficacy of immunotherapy, in our independent LUSC cohort and two validation sets, the ROC AUC of IPTS molecular typing was increased by 24%, 22% and 8% respectively compared with that of PD-L1 expression level. The results suggest that the prediction effect of our model is similar to that of TIDE or TIS. However, compared with TIDE, which needs to use whole gene transcriptome data to conduct online prediction, or TIS, which only knows the gene type and does not disclose the relevant calculation equations, and requires a special analysis system, building a IPTS model equation to predict the efficacy of immunotherapy have the advantages of lower cost and more convenience.
In our study, a total of 17 genes were screened to construct a predictive model for immunotherapy efficacy in patients with LUSC, of which 9 genes (AKAP2, GCGR, LRRC38, MARCO, NANOG, NTSR1, PF4, RP1, and TMEM236) were gene signatures of C1, and 8 genes (ALOX5, FCGR2A, KCNQ3, NLRP12, SCARF1, SIGLEC12, TGM2, and VSTM1) were gene signatures of C2. In previous studies, some of these genes have been associated with cancer progression and prognosis. Among these, AKAP2 was found to be upregulated in ovarian cancer, and promotes cancer cell growth as well as migration [55]. Increased expression of AKAP2 has been linked to metastatic prostate cancer, while knocking down its expression could significantly reduce the tumorigenicity and metastatic ability of prostate cancer cells [56]. GCGR was found to be an independent prognostic factor for OS in patients with NSCLC [57]. The protein encoded by MARCO is a member of the scavenger receptor family. It has been shown that targeting the scavenger receptor MARCO with antibodies reduces tumor growth and metastasis in murine tumor models of melanoma, colon cancer, and breast cancer [58]. Furthermore, the homeobox-domain transcription factor NANOG, a key regulator of embryonic development and cellular reprogramming, is ubiquitously expressed in human cancers [59]. Its overexpression has been linked to a worse prognosis in lung cancer [60]. NTSR1 is reportedly expressed in 40% of lung tumors, and its expression is a negative prognostic marker in patients with surgically resected stage I lung adenocarcinoma [61]. PF4 is a cancer-enhancing endocrine signal, and its overexpression in tumors is associated with reduced OS in patients with lung cancer [62]. As six of the nine genes associated with low immune cell infiltration (type C1) were involved in the pathogenesis, malignant transformation, and progression of a variety of cancers, including LUSC, as well as showing significant correlations with patient survival and prognosis, the findings of our bioinformatics analysis are meaningful to an extent.
Among the 8-gene signature of high immune cell infiltration (type C2), ALOX5 has been found to promote gastric cancer growth and attenuate chemotherapy toxicity [63], while in breast cancer, ALOX5 activation is associated with HER2 expression as well as mediates breast cancer growth and migration [64]. Recent studies have reported that the polymorphism of FCGR2A expression is associated with an increased risk of lung cancer [65]. NLRP12 is a key factor in maintaining intestinal homeostasis and preventing colorectal tumors [66]. Higher SCARF1 expression in hepatocellular carcinoma tumor tissues was highly prognostic of better OS, DFS and PFS [67]. High frequency of SIGLEC12 expression in advanced colorectal cancer cohort and correlation with OS [68]. TGM2 has been shown to enhance the migration and invasion of lung cancer cells [69]. TMEM236 has the potential to be a potential novel diagnostic biomarker for colorectal cancer [70]. Downregulated in bone marrow cells from leukemia patients, VSTM1 may become a diagnostic and treatment target [71]. Only the three remaining genes, KCNQ3, LRRC38 and RP1, were rarely reported in any cancer research, and thus show potential value for research in LUSC.
Among these 17 genes, 13 genes were reported to be associated with immune-related pathways. The pathway with the largest number of associated genes is the mitogen-activated protein kinases (MAPKs) signaling pathway. Eight genes could regulate it, and they are AKAP2 [72], ALOX5 [63], GCGR [73], NLRP12 [74], NTSR1 [75], PF4 [76], SIGLEC12 [68] and TGM2 [77]. Wnt/β-catenin signaling pathway could be regulated by AKAP2 [55], ALOX5 [78] and TGM2 [79]. PI3K/AKT/mammalian target of rapamycin (mTOR) signaling pathway could be regulated by ALOX5 [80], SCARF1 [81] and TGM2 [82]. There are seven genes involved in the regulation of nuclear transcription factor-κB (NF-κB) signaling pathway, such as ALOX5 [83], MARCO [84], NANOG [85], NLRP12 [74], NTSR1 [75], TGM2 [86], and VSTM1 [87]. Toll-like receptors (TLRs) signaling pathway could be regulated by FCGR2A [88], MARCO [89], NANOG [90], NLRP12 [74], and PF4 [91]. Six genes could involved in the regulation of janus kinase/signal transducer and activator of transcription (JAK/STAT) signaling pathway, such as ALOX5 [78, 92], MARCO [93], NLRP12 [94], PF4 [95], SCARF1 [81], and TGM2 [96]. In addition, some genes have other immune-related functions. For instance, ALOX5 contributes to the recruitment and activation of macrophages thereby adding to the role of macrophages in a dynamically changing tumor environment [97]. FCGR2A encodes the receptor protein on the surface of immune cells, which can transmit activation signals to cells through its tyrosine-based activation motif [98]. Antibodies targeting MARCO in NSCLC restore the anti-tumor activity of T cells and NK cells by polarizing suppressor macrophages [99]. NLRP12 plays critical roles in balancing T cell response to control overt activation and maintain cellular homeostasis [100]. SCARF1 mediates the clearance of apoptotic cells and prevents autoimmunity [101]. SIGLEC12 encodes one of the CD33-related SIGLEC family of signaling molecules in immune cells [102]. The binding of the TGM2 mediated crosslinked fibrinogens to un-stimulated endothelial cells can assemble leukocytes, platelets or fibrin, and promote inflammation [103]. Only the four remaining genes, KCNQ3, LRRC38, RP1 and TMEM236 were rarely reported in any immue-related research, which provides new ideas for follow-up studies based on these four genes, especially in the immunological research related to LUSC.
ICIs enhance T cell activity by blocking CTLA-4, PD-1, or PD-L1 to achieve an anti-tumor effect. The high expression of CTLA-4, PD-1, and PD-L1/PD-L2 has been positively correlated with the efficacy of immunotherapy, which has a certain value for therapeutic prediction [104]. By exploring the relationship between IPTS and the expression of CTLA-4, PD-1, PD-L1, and PD-L2, we found that the expression of four immuno-inhibitors was significantly positively correlated with the IPTS in the high score group. In addition, the difference analysis of immune molecular typing between the two IPTS subgroups (either high or low scores) revealed that the enrichment scores of chemokines, chemokine receptors, MHC molecules, immuno-inhibitors, and immuno-stimulators in patients with high IPTS were significantly higher than those in patients with low IPTS. These findings further indicated evident differences in the immune microenvironment between these two subtypes, with tumors in the high score group more likely to be “hot tumors”.
Our study found that patients with high IPTS had a worse prognosis than those with low IPTS in the training set (patients not receiving immunotherapy), while in the validation set GSE135222 and our LUSC cohort (patients receiving immunotherapy), this situation had been reversed. In other words, patients with high IPTS were more likely to benefit from immunotherapy than those with low IPTS. As for patients with low IPTS, we further explored the correlation between IPTS and anti-tumor drug efficacy, and found that the IC50 of five drugs (i.e., acetalax, AZD2014, GSK2606414, obatoclax mesylate, and VSP34_8731) in LUSC cells with high IPTS was higher than that in cells with low IPTS, suggesting that patients with low IPTS might be sensitive to these drugs. Among them, acetalax, also known as oxyphenisatin acetate, has shown antitumor activity in mouse xenograft models by inducing tumor necrosis factor (TNF) α expression and TNFR1 degradation, indicating autocrine TNF α-mediated apoptosis. AZD2014 is a mTOR inhibitor [105]. mTOR is a key kinase of PI3K/AKT/mTOR signaling pathway, which can regulate the tumor cell proliferation, differentiation, apoptosis and other processes. Previous studies have shown that mTOR signaling pathway has a significant regulatory effect on immune function and T cell differentiation by integrating various microenvironment signals [106, 107]. AZD2014 has been proved to have dramatic anti-tumor effects in phase II clinical trials for breast cancer [108] and hepatocellular carcinoma [109]. As a protein kinase R-like endoplasmic reticulum kinase (PERK) inhibitor, GSK2606414 can significantly inhibit the PERK dependent signaling pathway in human colorectal adenocarcinoma cell line HT-29 and human neuroblastoma cell lines SH-SY5Y, which can promote apoptosis by inducing endoplasmic reticulum stress [110, 111]. The pan-Bcl-2 inhibitor Obatoclax can sensitize hepatocellular carcinoma cells to promote the anti-tumor efficacy in combination with ICIs, for Obatoclax can sensitize T cell mediated killing by promoting T cell activation and the expression of effector cytokines in spleen and tumor [112]. VSP34, as a type III phosphatidylinositol kinase, is a key protein in the process of autophagy [113]. Recently, Noman et al. [114] reported that VSP34 regulated the TME through its kinase activity, and VSP34 protein knockdown or VSP34 kinase activity inhibition could transform tumors from “cold tumors” to “hot tumors” to enhance the effect of ICIs. As an inhibitor targeting VSP34, VSP34_8731 has the potential to realize the transition from C1 tumors to C2 tumors by increasing the infiltration of immune cells into tumor tissues. It can be concluded from the above studies that these five drugs have the effects of regulating immune process thereby promoting tumor cell apoptosis, and it might be the reason that the LUSC cell lines with low IPTSes may be more sensitive to these five antitumor drugs. This also demonstrates the feasibility of our study in using high and low immune cell infiltration typing for patients with LUSC as a measure of immunotherapy efficacy, and our findings provided a theoretical basis for the selection of treatment methods in patients with LUSC, and also put forth a new treatment scheme with potential curative effect for patients with poor outcomes after immunotherapy.
Our study presented a potential new method for predicting the efficacy of immunotherapy in LUSC. Nevertheless, there are still some limitations that should not be ignored. First, based on the data from public databases, the internal mechanism still needs experimental verification. Through functional enrichment analysis, it was found that the high IPTS groups involved the regulation of multiple pathways related to tumor occurrence and development, which requires follow-up molecular mechanism research. Second, due to the different sequencing platforms of the training set (TCGA-LUSC) and validation sets (GSE126044 and GSE135222) giving rise to different sequencing backgrounds and normalization methods, it is difficult to obtain the best IPTS value suitable for all data sets to distinguish high or low immune cells infiltration. Therefore, the initial IPTS threshold should be obtained through small sample testing, and then corrected by conducting a large-scale prospective clinical study. Furthermore, whether a high IPTS could become a predictor of immunotherapy efficacy also needs to be further confirmed by large-scale prospective clinical trials. Third, regarding anti-tumor drug treatment, the number of LUSC cell lines in the GDSC database is relatively small at only 15. To maximize the test efficiency, we grouped them as high and low IPTS groups according to 1:1; hence, there is likely to be a certain bias. The results of this study may still provide theoretical support for the treatment of LUSC with anti-tumor drugs.

Conclusion

In conclusion, we constructed a model containing 17 genes to predict the efficacy of immunotherapy for patients with LUSC based on bioinformatics analysis on the TCGA database. The prediction effect of the model was verified in two independent cohorts in the GEO database. The IPTS molecular typing positively correlated with both the degree of tumor immune cell infiltration and the efficacy of immunotherapy with potential prognostic value. This study provides a new method for predicting the efficacy of immunotherapy for LUSC, which may have potential clinical prospects.

Acknowledgements

We thank those authors who released and shared their datasets on the TCGA and GEO databases. We are grateful to Sangerbox (http://​vip.​sangerbox.​com) for providing technical support and Bullet Edits Limited for copyediting the manuscript.

Declarations

The studies involving human participants were reviewed and approved by the Medical Ethics Committee of the Second Affiliated Hospital of Zhejiang University School of Medicine (Approval Number: 2022-0548/I2022685). The patients/participants provided their written informed consent to participate in this study.
Not applicable.

Competing interests

The authors declare that they have no competing interests.
Open AccessThis article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​. The Creative Commons Public Domain Dedication waiver (http://​creativecommons.​org/​publicdomain/​zero/​1.​0/​) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
insite
INHALT
download
DOWNLOAD
print
DRUCKEN
Literatur
1.
Zurück zum Zitat Sung H, Ferlay J, Siegel RL, Laversanne M, Soerjomataram I, Jemal A, et al. global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J Clin. 2021;71(3):209–49.PubMedCrossRef Sung H, Ferlay J, Siegel RL, Laversanne M, Soerjomataram I, Jemal A, et al. global cancer statistics 2020: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J Clin. 2021;71(3):209–49.PubMedCrossRef
2.
Zurück zum Zitat Perez-Moreno P, Brambilla E, Thomas R, Soria J. Squamous cell carcinoma of the lung: molecular subtypes and therapeutic opportunities. Clin Cancer Res. 2012;18(9):2443–51.PubMedCrossRef Perez-Moreno P, Brambilla E, Thomas R, Soria J. Squamous cell carcinoma of the lung: molecular subtypes and therapeutic opportunities. Clin Cancer Res. 2012;18(9):2443–51.PubMedCrossRef
3.
Zurück zum Zitat Zhou H, Zhang H, Shi M, Wang J, Huang Z, Shi J. A robust signature associated with patient prognosis and tumor immune microenvironment based on immune-related genes in lung squamous cell carcinoma. Int Immunopharmacol. 2020;88: 106856.PubMedCrossRef Zhou H, Zhang H, Shi M, Wang J, Huang Z, Shi J. A robust signature associated with patient prognosis and tumor immune microenvironment based on immune-related genes in lung squamous cell carcinoma. Int Immunopharmacol. 2020;88: 106856.PubMedCrossRef
4.
Zurück zum Zitat Thorsteinsson H, Alexandersson A, Oskarsdottir GN, Skuladottir R, Isaksson HJ, Jonsson S, et al. Resection rate and outcome of pulmonary resections for non-small-cell lung cancer: a nationwide study from Iceland. J Thorac Oncol. 2012;7(7):1164–9.PubMedCrossRef Thorsteinsson H, Alexandersson A, Oskarsdottir GN, Skuladottir R, Isaksson HJ, Jonsson S, et al. Resection rate and outcome of pulmonary resections for non-small-cell lung cancer: a nationwide study from Iceland. J Thorac Oncol. 2012;7(7):1164–9.PubMedCrossRef
5.
Zurück zum Zitat Cancer Genome Atlas Research Network. Comprehensive genomic characterization of squamous cell lung cancers. Nature. 2012;489(7417):519–25.CrossRef Cancer Genome Atlas Research Network. Comprehensive genomic characterization of squamous cell lung cancers. Nature. 2012;489(7417):519–25.CrossRef
6.
Zurück zum Zitat Kris MG, Johnson BE, Berry LD, Kwiatkowski DJ, Iafrate AJ, Wistuba II, et al. Using multiplexed assays of oncogenic drivers in lung cancers to select targeted drugs. JAMA. 2014;311(19):1998.PubMedPubMedCentralCrossRef Kris MG, Johnson BE, Berry LD, Kwiatkowski DJ, Iafrate AJ, Wistuba II, et al. Using multiplexed assays of oncogenic drivers in lung cancers to select targeted drugs. JAMA. 2014;311(19):1998.PubMedPubMedCentralCrossRef
7.
Zurück zum Zitat Shtivelman E, Hensing T, Simon GR, Dennis PA, Otterson GA, Bueno R, et al. Molecular pathways and therapeutic targets in lung cancer. Oncotarget. 2014;5(6):1392–433.PubMedPubMedCentralCrossRef Shtivelman E, Hensing T, Simon GR, Dennis PA, Otterson GA, Bueno R, et al. Molecular pathways and therapeutic targets in lung cancer. Oncotarget. 2014;5(6):1392–433.PubMedPubMedCentralCrossRef
8.
Zurück zum Zitat Song X, Zhao C, Jiang L, Lin S, Bi J, Wei Q, et al. High PITX1 expression in lung adenocarcinoma patients is associated with DNA methylation and poor prognosis. Pathol Res Pract. 2018;214(12):2046–53.PubMedCrossRef Song X, Zhao C, Jiang L, Lin S, Bi J, Wei Q, et al. High PITX1 expression in lung adenocarcinoma patients is associated with DNA methylation and poor prognosis. Pathol Res Pract. 2018;214(12):2046–53.PubMedCrossRef
9.
Zurück zum Zitat Paik PK, Pillai RN, Lathan CS, Velasco SA, Papadimitrakopoulou V. New treatment options in advanced squamous cell lung cancer. Am Soc Clin Oncol Educ Book. 2019;39(39):e198-206.PubMedCrossRef Paik PK, Pillai RN, Lathan CS, Velasco SA, Papadimitrakopoulou V. New treatment options in advanced squamous cell lung cancer. Am Soc Clin Oncol Educ Book. 2019;39(39):e198-206.PubMedCrossRef
10.
Zurück zum Zitat Darvin P, Toor SM, Sasidharan Nair V, Elkord E. Immune checkpoint inhibitors: recent progress and potential biomarkers. Exp Mol Med. 2018;50(12):1–11.PubMedCrossRef Darvin P, Toor SM, Sasidharan Nair V, Elkord E. Immune checkpoint inhibitors: recent progress and potential biomarkers. Exp Mol Med. 2018;50(12):1–11.PubMedCrossRef
11.
Zurück zum Zitat Doroshow DB, Sanmamed MF, Hastings K, Politi K, Rimm DL, Chen L, et al. Immunotherapy in non-small cell lung cancer: facts and hopes. Clin Cancer Res. 2019;25(15):4592–602.PubMedPubMedCentralCrossRef Doroshow DB, Sanmamed MF, Hastings K, Politi K, Rimm DL, Chen L, et al. Immunotherapy in non-small cell lung cancer: facts and hopes. Clin Cancer Res. 2019;25(15):4592–602.PubMedPubMedCentralCrossRef
12.
Zurück zum Zitat Yang F, Wang JF, Wang Y, Liu B, Molina JR. Comparative analysis of predictive biomarkers for PD-1/PD-L1 inhibitors in cancers: developments and challenges. Cancers. 2022;14(1):109.CrossRef Yang F, Wang JF, Wang Y, Liu B, Molina JR. Comparative analysis of predictive biomarkers for PD-1/PD-L1 inhibitors in cancers: developments and challenges. Cancers. 2022;14(1):109.CrossRef
13.
Zurück zum Zitat Riaz N, Havel JJ, Makarov V, Desrichard A, Urba WJ, Sims JS, et al. Tumor and microenvironment evolution during immunotherapy with nivolumab. Cell. 2017;171(4):934–49.PubMedPubMedCentralCrossRef Riaz N, Havel JJ, Makarov V, Desrichard A, Urba WJ, Sims JS, et al. Tumor and microenvironment evolution during immunotherapy with nivolumab. Cell. 2017;171(4):934–49.PubMedPubMedCentralCrossRef
15.
Zurück zum Zitat Wang Y, Lin X, Sun D. A narrative review of prognosis prediction models for non-small cell lung cancer: what kind of predictors should be selected and how to improve models? Ann Transl Med. 2021;9(20):1597.PubMedPubMedCentralCrossRef Wang Y, Lin X, Sun D. A narrative review of prognosis prediction models for non-small cell lung cancer: what kind of predictors should be selected and how to improve models? Ann Transl Med. 2021;9(20):1597.PubMedPubMedCentralCrossRef
16.
Zurück zum Zitat Wu Y, Yang L, Zhang L, Zheng X, Xu H, Wang K, et al. Identification of a four-gene signature associated with the prognosis prediction of lung adenocarcinoma based on integrated bioinformatics analysis. Genes-Basel. 2022;13(2):238.PubMedPubMedCentralCrossRef Wu Y, Yang L, Zhang L, Zheng X, Xu H, Wang K, et al. Identification of a four-gene signature associated with the prognosis prediction of lung adenocarcinoma based on integrated bioinformatics analysis. Genes-Basel. 2022;13(2):238.PubMedPubMedCentralCrossRef
17.
Zurück zum Zitat Yang L, Wu Y, Xu H, Zhang J, Zheng X, Zhang L, et al. Identification and validation of a novel six-lncRNA-based prognostic model for lung adenocarcinoma. Front Oncol. 2022;11: 775583.PubMedPubMedCentralCrossRef Yang L, Wu Y, Xu H, Zhang J, Zheng X, Zhang L, et al. Identification and validation of a novel six-lncRNA-based prognostic model for lung adenocarcinoma. Front Oncol. 2022;11: 775583.PubMedPubMedCentralCrossRef
18.
Zurück zum Zitat Gaujoux R, Seoighe C. A flexible R package for nonnegative matrix factorization. BMC Bioinform. 2010;11(1):367.CrossRef Gaujoux R, Seoighe C. A flexible R package for nonnegative matrix factorization. BMC Bioinform. 2010;11(1):367.CrossRef
19.
Zurück zum Zitat Tibshirani R. Regression shrinkage and selection via the lasso. J R Stat Soc Ser B (Methodol). 1996;58(1):267–88. Tibshirani R. Regression shrinkage and selection via the lasso. J R Stat Soc Ser B (Methodol). 1996;58(1):267–88.
20.
Zurück zum Zitat Frankish A, Diekhans M, Jungreis I, Lagarde J, Loveland JE, Mudge JM, et al. Gencode 2021. Nucleic Acids Res. 2021;49(D1):D916–23.PubMedCrossRef Frankish A, Diekhans M, Jungreis I, Lagarde J, Loveland JE, Mudge JM, et al. Gencode 2021. Nucleic Acids Res. 2021;49(D1):D916–23.PubMedCrossRef
21.
Zurück zum Zitat Cho J, Hong MH, Ha S, Kim Y, Cho BC, Lee I, et al. Genome-wide identification of differentially methylated promoters and enhancers associated with response to anti-PD-1 therapy in non-small cell lung cancer. Exp Mol Med. 2020;52(9):1550–63.PubMedPubMedCentralCrossRef Cho J, Hong MH, Ha S, Kim Y, Cho BC, Lee I, et al. Genome-wide identification of differentially methylated promoters and enhancers associated with response to anti-PD-1 therapy in non-small cell lung cancer. Exp Mol Med. 2020;52(9):1550–63.PubMedPubMedCentralCrossRef
22.
Zurück zum Zitat Jung H, Kim HS, Kim JY, Sun J, Ahn JS, Ahn M, et al. DNA methylation loss promotes immune evasion of tumours with high mutation and copy number load. Nat Commun. 2019;10(1):4278.PubMedPubMedCentralCrossRef Jung H, Kim HS, Kim JY, Sun J, Ahn JS, Ahn M, et al. DNA methylation loss promotes immune evasion of tumours with high mutation and copy number load. Nat Commun. 2019;10(1):4278.PubMedPubMedCentralCrossRef
23.
Zurück zum Zitat Kim JY, Choi JK, Jung H. Genome-wide methylation patterns predict clinical benefit of immunotherapy in lung cancer. Clin Epigenet. 2020;12(1):119.CrossRef Kim JY, Choi JK, Jung H. Genome-wide methylation patterns predict clinical benefit of immunotherapy in lung cancer. Clin Epigenet. 2020;12(1):119.CrossRef
24.
Zurück zum Zitat Jia Q, Wu W, Wang Y, Alexander PB, Sun C, Gong Z, et al. Local mutational diversity drives intratumoral immune heterogeneity in non-small cell lung cancer. Nat Commun. 2018;9(1):5361.PubMedPubMedCentralCrossRef Jia Q, Wu W, Wang Y, Alexander PB, Sun C, Gong Z, et al. Local mutational diversity drives intratumoral immune heterogeneity in non-small cell lung cancer. Nat Commun. 2018;9(1):5361.PubMedPubMedCentralCrossRef
25.
Zurück zum Zitat Hänzelmann S, Castelo R, Guinney J. GSVA: gene set variation analysis for microarray and RNA-seq data. BMC Bioinform. 2013;14(1):7.CrossRef Hänzelmann S, Castelo R, Guinney J. GSVA: gene set variation analysis for microarray and RNA-seq data. BMC Bioinform. 2013;14(1):7.CrossRef
26.
Zurück zum Zitat Gaujoux R, Seoighe C. A flexible R package for nonnegative matrix factorization. BMC Bioinform. 2010;11:367.CrossRef Gaujoux R, Seoighe C. A flexible R package for nonnegative matrix factorization. BMC Bioinform. 2010;11:367.CrossRef
27.
Zurück zum Zitat Krijthe JH. Rtsne: T-distributed stochastic neighbor embedding using barnes-hut implementation. R package version 0.15. 2015. Krijthe JH. Rtsne: T-distributed stochastic neighbor embedding using barnes-hut implementation. R package version 0.15. 2015.
28.
Zurück zum Zitat R Core Team. R: a language and environment for statistical computing. Vienna: R Foundation for Statistical Computing; 2021. R Core Team. R: a language and environment for statistical computing. Vienna: R Foundation for Statistical Computing; 2021.
29.
Zurück zum Zitat Ritchie ME, Phipson B, Wu D, Hu Y, Law CW, Shi W, et al. Limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res. 2015;43(7): e47.PubMedPubMedCentralCrossRef Ritchie ME, Phipson B, Wu D, Hu Y, Law CW, Shi W, et al. Limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res. 2015;43(7): e47.PubMedPubMedCentralCrossRef
30.
Zurück zum Zitat Wu T, Hu E, Xu S, Chen M, Guo P, Dai Z, et al. ClusterProfiler 4.0: a universal enrichment tool for interpreting omics data. Innovation. 2021;2(3): 100141.PubMedPubMedCentral Wu T, Hu E, Xu S, Chen M, Guo P, Dai Z, et al. ClusterProfiler 4.0: a universal enrichment tool for interpreting omics data. Innovation. 2021;2(3): 100141.PubMedPubMedCentral
31.
Zurück zum Zitat Therneau T. A package for survival analysis in R. R package version 3.3-1. 2022. Therneau T. A package for survival analysis in R. R package version 3.3-1. 2022.
32.
Zurück zum Zitat Alboukadel Kassambara MKAP. Survminer: drawing survival curves using 'ggplot2'. R package version 0.4.9. 2021. Alboukadel Kassambara MKAP. Survminer: drawing survival curves using 'ggplot2'. R package version 0.4.9. 2021.
33.
34.
Zurück zum Zitat Simon N, Friedman J, Hastie T, Tibshirani R. Regularization paths for Cox’s proportional hazards model via coordinate descent. J Stat Softw. 2011;39(5):1–13.PubMedPubMedCentralCrossRef Simon N, Friedman J, Hastie T, Tibshirani R. Regularization paths for Cox’s proportional hazards model via coordinate descent. J Stat Softw. 2011;39(5):1–13.PubMedPubMedCentralCrossRef
35.
Zurück zum Zitat Frank E Harrell Jr. rms: Regression Modeling Strategies. R package version 6.2-0. 2021. Frank E Harrell Jr. rms: Regression Modeling Strategies. R package version 6.2-0. 2021.
36.
Zurück zum Zitat J.J. Allaire, Christopher Gandrud, Kenton Russell, CJ Yetman. networkD3: D3 JavaScript network graphs from R. R package version 0.4. 2017. J.J. Allaire, Christopher Gandrud, Kenton Russell, CJ Yetman. networkD3: D3 JavaScript network graphs from R. R package version 0.4. 2017.
37.
Zurück zum Zitat Robin X, Turck N, Hainard A, Tiberti N, Lisacek F, Sanchez J, et al. pROC: an open-source package for R and S+ to analyze and compare ROC curves. BMC Bioinform. 2011;12(1):77.CrossRef Robin X, Turck N, Hainard A, Tiberti N, Lisacek F, Sanchez J, et al. pROC: an open-source package for R and S+ to analyze and compare ROC curves. BMC Bioinform. 2011;12(1):77.CrossRef
38.
Zurück zum Zitat Yang W, Soares J, Greninger P, Edelman EJ, Lightfoot H, Forbes S, et al. Genomics of drug sensitivity in cancer (GDSC): a resource for therapeutic biomarker discovery in cancer cells. Nucleic Acids Res. 2012;41(D1):D955–61.PubMedPubMedCentralCrossRef Yang W, Soares J, Greninger P, Edelman EJ, Lightfoot H, Forbes S, et al. Genomics of drug sensitivity in cancer (GDSC): a resource for therapeutic biomarker discovery in cancer cells. Nucleic Acids Res. 2012;41(D1):D955–61.PubMedPubMedCentralCrossRef
39.
Zurück zum Zitat Yoshihara K, Shahmoradgoli M, Martínez E, Vegesna R, Kim H, Torres-Garcia W, et al. Inferring tumour purity and stromal and immune cell admixture from expression data. Nat Commun. 2013;4(1):2612.PubMedCrossRef Yoshihara K, Shahmoradgoli M, Martínez E, Vegesna R, Kim H, Torres-Garcia W, et al. Inferring tumour purity and stromal and immune cell admixture from expression data. Nat Commun. 2013;4(1):2612.PubMedCrossRef
40.
Zurück zum Zitat Charoentong P, Finotello F, Angelova M, Mayer C, Efremova M, Rieder D, et al. Pan-cancer immunogenomic analyses reveal genotype-immunophenotype relationships and predictors of response to checkpoint blockade. Cell Rep. 2017;18(1):248–62.PubMedCrossRef Charoentong P, Finotello F, Angelova M, Mayer C, Efremova M, Rieder D, et al. Pan-cancer immunogenomic analyses reveal genotype-immunophenotype relationships and predictors of response to checkpoint blockade. Cell Rep. 2017;18(1):248–62.PubMedCrossRef
41.
Zurück zum Zitat Hu J, Yu A, Othmane B, Qiu D, Li H, Li C, et al. Siglec15 shapes a non-inflamed tumor microenvironment and predicts the molecular subtype in bladder cancer. Theranostics. 2021;11(7):3089–108.PubMedPubMedCentralCrossRef Hu J, Yu A, Othmane B, Qiu D, Li H, Li C, et al. Siglec15 shapes a non-inflamed tumor microenvironment and predicts the molecular subtype in bladder cancer. Theranostics. 2021;11(7):3089–108.PubMedPubMedCentralCrossRef
42.
Zurück zum Zitat Subramanian A, Tamayo P, Mootha VK, Mukherjee S, Ebert BL, Gillette MA, et al. Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc Natl Acad Sci. 2005;102(43):15545–50.PubMedPubMedCentralCrossRef Subramanian A, Tamayo P, Mootha VK, Mukherjee S, Ebert BL, Gillette MA, et al. Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc Natl Acad Sci. 2005;102(43):15545–50.PubMedPubMedCentralCrossRef
43.
Zurück zum Zitat Wickham H. ggplot2: elegant graphics for data analysis. New York: Springer-Verlag; 2016.CrossRef Wickham H. ggplot2: elegant graphics for data analysis. New York: Springer-Verlag; 2016.CrossRef
44.
Zurück zum Zitat Allemani C, Matsuda T, Di Carlo V, Harewood R, Matz M, Niksic M, et al. Global surveillance of trends in cancer survival 2000–14 (CONCORD-3): analysis of individual records for 37 513 025 patients diagnosed with one of 18 cancers from 322 population-based registries in 71 countries. Lancet. 2018;391(10125):1023–75.PubMedPubMedCentralCrossRef Allemani C, Matsuda T, Di Carlo V, Harewood R, Matz M, Niksic M, et al. Global surveillance of trends in cancer survival 2000–14 (CONCORD-3): analysis of individual records for 37 513 025 patients diagnosed with one of 18 cancers from 322 population-based registries in 71 countries. Lancet. 2018;391(10125):1023–75.PubMedPubMedCentralCrossRef
45.
Zurück zum Zitat Francisci S, Minicozzi P, Pierannunzio D, Ardanaz E, Eberle A, Grimsrud TK, et al. Survival patterns in lung and pleural cancer in Europe 1999–2007: results from the EUROCARE-5 study. Eur J Cancer. 2015;51(15):2242–53.PubMedCrossRef Francisci S, Minicozzi P, Pierannunzio D, Ardanaz E, Eberle A, Grimsrud TK, et al. Survival patterns in lung and pleural cancer in Europe 1999–2007: results from the EUROCARE-5 study. Eur J Cancer. 2015;51(15):2242–53.PubMedCrossRef
46.
Zurück zum Zitat Garon EB, Hellmann MD, Rizvi NA, Carcereny E, Leighl NB, Ahn M, et al. Five-year overall survival for patients with advanced non-small-cell lung cancer treated with pembrolizumab: results from the phase I KEYNOTE-001 study. J Clin Oncol. 2019;37(28):2518–27.PubMedPubMedCentralCrossRef Garon EB, Hellmann MD, Rizvi NA, Carcereny E, Leighl NB, Ahn M, et al. Five-year overall survival for patients with advanced non-small-cell lung cancer treated with pembrolizumab: results from the phase I KEYNOTE-001 study. J Clin Oncol. 2019;37(28):2518–27.PubMedPubMedCentralCrossRef
47.
Zurück zum Zitat Gettinger S, Horn L, Jackman D, Spigel D, Antonia S, Hellmann M, et al. Five-year follow-up of nivolumab in previously treated advanced non-small-cell lung cancer: results from the CA209-003 study. J Clin Oncol. 2018;36(17):1675–84.PubMedCrossRef Gettinger S, Horn L, Jackman D, Spigel D, Antonia S, Hellmann M, et al. Five-year follow-up of nivolumab in previously treated advanced non-small-cell lung cancer: results from the CA209-003 study. J Clin Oncol. 2018;36(17):1675–84.PubMedCrossRef
48.
Zurück zum Zitat Topalian SL, Hodi FS, Brahmer JR, Gettinger SN, Smith DC, Mcdermott DF, et al. Five-year survival and correlates among patients with advanced melanoma, renal cell carcinoma, or non-small cell lung cancer treated with nivolumab. JAMA Oncol. 2019;5(10):1411.PubMedPubMedCentralCrossRef Topalian SL, Hodi FS, Brahmer JR, Gettinger SN, Smith DC, Mcdermott DF, et al. Five-year survival and correlates among patients with advanced melanoma, renal cell carcinoma, or non-small cell lung cancer treated with nivolumab. JAMA Oncol. 2019;5(10):1411.PubMedPubMedCentralCrossRef
49.
Zurück zum Zitat Malhotra J, Jabbour SK, Aisner J. Current state of immunotherapy for non-small cell lung cancer. Transl Lung Cancer Res. 2007;6(2):196–211.CrossRef Malhotra J, Jabbour SK, Aisner J. Current state of immunotherapy for non-small cell lung cancer. Transl Lung Cancer Res. 2007;6(2):196–211.CrossRef
50.
Zurück zum Zitat Tian Y, Wang J, Wen Q, Su G, Sun Y. Immune subgroup analysis for non-small cell lung cancer may be a good choice for evaluating therapeutic efficacy and prognosis. Aging. 2021;13(9):12691–709.PubMedPubMedCentralCrossRef Tian Y, Wang J, Wen Q, Su G, Sun Y. Immune subgroup analysis for non-small cell lung cancer may be a good choice for evaluating therapeutic efficacy and prognosis. Aging. 2021;13(9):12691–709.PubMedPubMedCentralCrossRef
51.
Zurück zum Zitat Garon EB, Rizvi NA, Hui R, Leighl N, Balmanoukian AS, Eder JP, et al. Pembrolizumab for the treatment of non-small-cell lung cancer. New Engl J Med. 2015;372(21):2018–28.PubMedCrossRef Garon EB, Rizvi NA, Hui R, Leighl N, Balmanoukian AS, Eder JP, et al. Pembrolizumab for the treatment of non-small-cell lung cancer. New Engl J Med. 2015;372(21):2018–28.PubMedCrossRef
52.
Zurück zum Zitat Jiang P, Gu S, Pan D, Fu J, Sahu A, Hu X, et al. Signatures of T cell dysfunction and exclusion predict cancer immunotherapy response. Nat Med. 2018;24(10):1550–8.PubMedPubMedCentralCrossRef Jiang P, Gu S, Pan D, Fu J, Sahu A, Hu X, et al. Signatures of T cell dysfunction and exclusion predict cancer immunotherapy response. Nat Med. 2018;24(10):1550–8.PubMedPubMedCentralCrossRef
53.
Zurück zum Zitat Ayers M, Lunceford J, Nebozhyn M, Murphy E, Loboda A, Kaufman DR, et al. IFN-γ-related mRNA profile predicts clinical response to PD-1 blockade. J Clin Invest. 2017;127(8):2930–40.PubMedPubMedCentralCrossRef Ayers M, Lunceford J, Nebozhyn M, Murphy E, Loboda A, Kaufman DR, et al. IFN-γ-related mRNA profile predicts clinical response to PD-1 blockade. J Clin Invest. 2017;127(8):2930–40.PubMedPubMedCentralCrossRef
54.
Zurück zum Zitat Danaher P, Warren S, Lu R, Samayoa J, Sullivan A, Pekker I, et al. Pan-cancer adaptive immune resistance as defined by the tumor inflammation signature (TIS): results from the Cancer Genome Atlas (TCGA). J Immunother Cancer. 2018;6(1):63.PubMedPubMedCentralCrossRef Danaher P, Warren S, Lu R, Samayoa J, Sullivan A, Pekker I, et al. Pan-cancer adaptive immune resistance as defined by the tumor inflammation signature (TIS): results from the Cancer Genome Atlas (TCGA). J Immunother Cancer. 2018;6(1):63.PubMedPubMedCentralCrossRef
55.
Zurück zum Zitat Li X, Wang C, Zhang G, Liang M, Zhang B. AKAP2 is upregulated in ovarian cancer, and promotes growth and migration of cancer cells. Mol Med Rep. 2017;16(4):5151–6.PubMedPubMedCentralCrossRef Li X, Wang C, Zhang G, Liang M, Zhang B. AKAP2 is upregulated in ovarian cancer, and promotes growth and migration of cancer cells. Mol Med Rep. 2017;16(4):5151–6.PubMedPubMedCentralCrossRef
56.
Zurück zum Zitat Thakkar A, Aljameeli A, Thomas S, Shah GV. A-kinase anchoring protein 2 is required for calcitonin-mediated invasion of cancer cells. Endocr-Relat Cancer. 2016;23(1):1–14.PubMedCrossRef Thakkar A, Aljameeli A, Thomas S, Shah GV. A-kinase anchoring protein 2 is required for calcitonin-mediated invasion of cancer cells. Endocr-Relat Cancer. 2016;23(1):1–14.PubMedCrossRef
57.
Zurück zum Zitat Li R, Liu X, Zhou XJ, Chen X, Li JP, Yin YH, et al. Identification and validation of the prognostic value of immune-related genes in non-small cell lung cancer. Am J Transl Res. 2020;12(9):5844–65.PubMedPubMedCentral Li R, Liu X, Zhou XJ, Chen X, Li JP, Yin YH, et al. Identification and validation of the prognostic value of immune-related genes in non-small cell lung cancer. Am J Transl Res. 2020;12(9):5844–65.PubMedPubMedCentral
58.
Zurück zum Zitat Georgoudaki A, Prokopec KE, Boura VF, Hellqvist E, Sohn S, Östling J, et al. Reprogramming tumor-associated macrophages by antibody targeting inhibits cancer progression and metastasis. Cell Rep. 2016;15(9):2000–11.PubMedCrossRef Georgoudaki A, Prokopec KE, Boura VF, Hellqvist E, Sohn S, Östling J, et al. Reprogramming tumor-associated macrophages by antibody targeting inhibits cancer progression and metastasis. Cell Rep. 2016;15(9):2000–11.PubMedCrossRef
59.
Zurück zum Zitat Jeter CR, Yang T, Wang J, Chao H, Tang DG. Concise review: NANOG in cancer stem cells and tumor development: an update and outstanding questions. Stem Cells. 2015;33(8):2381–90.PubMedCrossRef Jeter CR, Yang T, Wang J, Chao H, Tang DG. Concise review: NANOG in cancer stem cells and tumor development: an update and outstanding questions. Stem Cells. 2015;33(8):2381–90.PubMedCrossRef
60.
Zurück zum Zitat Du Y, Ma C, Wang Z, Liu Z, Liu H, Wang T. Nanog, a novel prognostic marker for lung cancer. Surg Oncol. 2013;22(4):224–9.PubMedCrossRef Du Y, Ma C, Wang Z, Liu Z, Liu H, Wang T. Nanog, a novel prognostic marker for lung cancer. Surg Oncol. 2013;22(4):224–9.PubMedCrossRef
61.
Zurück zum Zitat Alifano M, Souazé F, Dupouy S, Camilleri-Broët S, Younes M, Ahmed-Zaïd S, et al. Neurotensin receptor 1 determines the outcome of non-small cell lung cancer. Clin Cancer Res. 2010;16(17):4401–10.PubMedCrossRef Alifano M, Souazé F, Dupouy S, Camilleri-Broët S, Younes M, Ahmed-Zaïd S, et al. Neurotensin receptor 1 determines the outcome of non-small cell lung cancer. Clin Cancer Res. 2010;16(17):4401–10.PubMedCrossRef
62.
Zurück zum Zitat Pucci F, Rickelt S, Newton AP, Garris C, Nunes E, Evavold C, et al. PF4 promotes platelet production and lung cancer growth. Cell Rep. 2016;17(7):1764–72.PubMedPubMedCentralCrossRef Pucci F, Rickelt S, Newton AP, Garris C, Nunes E, Evavold C, et al. PF4 promotes platelet production and lung cancer growth. Cell Rep. 2016;17(7):1764–72.PubMedPubMedCentralCrossRef
63.
Zurück zum Zitat Tang J, Zhang C, Lin J, Duan P, Long J, Zhu H. ALOX5-5-HETE promotes gastric cancer growth and alleviates chemotherapy toxicity via MEK/ERK activation. Cancer Med. 2021;10(15):5246–55.PubMedPubMedCentralCrossRef Tang J, Zhang C, Lin J, Duan P, Long J, Zhu H. ALOX5-5-HETE promotes gastric cancer growth and alleviates chemotherapy toxicity via MEK/ERK activation. Cancer Med. 2021;10(15):5246–55.PubMedPubMedCentralCrossRef
64.
Zurück zum Zitat Zhou X, Jiang Y, Li Q, Huang Z, Yang H, Wei C. Aberrant ALOX5 activation correlates with HER2 status and mediates breast cancer biological activities through multiple mechanisms. Biomed Res Int. 2020;2020:1–8. Zhou X, Jiang Y, Li Q, Huang Z, Yang H, Wei C. Aberrant ALOX5 activation correlates with HER2 status and mediates breast cancer biological activities through multiple mechanisms. Biomed Res Int. 2020;2020:1–8.
65.
Zurück zum Zitat He J, Yu L, Qiao Z, Yu B, Liu Y, Ren H. Genetic polymorphisms of FCGR2A, ORAI1 and CD40 are associated with risk of lung cancer. Eur J Cancer Prev. 2022;31(1):7–13.PubMedCrossRef He J, Yu L, Qiao Z, Yu B, Liu Y, Ren H. Genetic polymorphisms of FCGR2A, ORAI1 and CD40 are associated with risk of lung cancer. Eur J Cancer Prev. 2022;31(1):7–13.PubMedCrossRef
66.
Zurück zum Zitat Zaki MH, Vogel P, Malireddi RKS, Body-Malapel M, Anand PK, Bertin J, et al. The NOD-like receptor NLRP12 attenuates colon inflammation and tumorigenesis. Cancer Cell. 2011;20(5):649–60.PubMedPubMedCentralCrossRef Zaki MH, Vogel P, Malireddi RKS, Body-Malapel M, Anand PK, Bertin J, et al. The NOD-like receptor NLRP12 attenuates colon inflammation and tumorigenesis. Cancer Cell. 2011;20(5):649–60.PubMedPubMedCentralCrossRef
67.
Zurück zum Zitat Patten DA, Wilkinson AL, O’Rourke JM, Shetty S. Prognostic value and potential immunoregulatory role of SCARF1 in hepatocellular carcinoma. Front Oncol. 2020;10: 565950.PubMedPubMedCentralCrossRef Patten DA, Wilkinson AL, O’Rourke JM, Shetty S. Prognostic value and potential immunoregulatory role of SCARF1 in hepatocellular carcinoma. Front Oncol. 2020;10: 565950.PubMedPubMedCentralCrossRef
68.
Zurück zum Zitat Siddiqui SS, Vaill M, Do R, Khan N, Verhagen AL, Zhang W, et al. Human-specific polymorphic pseudogenization of SIGLEC12 protects against advanced cancer progression. FASEB Bioadv. 2021;3(2):69–82.PubMedCrossRef Siddiqui SS, Vaill M, Do R, Khan N, Verhagen AL, Zhang W, et al. Human-specific polymorphic pseudogenization of SIGLEC12 protects against advanced cancer progression. FASEB Bioadv. 2021;3(2):69–82.PubMedCrossRef
69.
Zurück zum Zitat Lee H, Huang C, Chen W, Tsai C, Chao Y, Liu S, et al. Transglutaminase 2 promotes migration and invasion of lung cancer cells. Oncol Res Featur Preclin Clin Cancer Ther. 2018;26(8):1175–82. Lee H, Huang C, Chen W, Tsai C, Chao Y, Liu S, et al. Transglutaminase 2 promotes migration and invasion of lung cancer cells. Oncol Res Featur Preclin Clin Cancer Ther. 2018;26(8):1175–82.
70.
Zurück zum Zitat Maurya NS, Kushwaha S, Chawade A, Mani A. Transcriptome profiling by combined machine learning and statistical R analysis identifies TMEM236 as a potential novel diagnostic biomarker for colorectal cancer. Sci Rep. 2021;11(1):14304.PubMedPubMedCentralCrossRef Maurya NS, Kushwaha S, Chawade A, Mani A. Transcriptome profiling by combined machine learning and statistical R analysis identifies TMEM236 as a potential novel diagnostic biomarker for colorectal cancer. Sci Rep. 2021;11(1):14304.PubMedPubMedCentralCrossRef
71.
Zurück zum Zitat Zhou J, Yao QM, Li JL, Chang Y, Li T, Han WL, et al. Synergistic antitumor activity of triple-regulated oncolytic adenovirus with VSTM1 and daunorubicin in leukemic cells. Apoptosis. 2016;21(10):1179–90.PubMedCrossRef Zhou J, Yao QM, Li JL, Chang Y, Li T, Han WL, et al. Synergistic antitumor activity of triple-regulated oncolytic adenovirus with VSTM1 and daunorubicin in leukemic cells. Apoptosis. 2016;21(10):1179–90.PubMedCrossRef
72.
Zurück zum Zitat Wang B, Jiang B, Li Y, Dai Y, Li P, Li L, et al. AKAP2 overexpression modulates growth plate chondrocyte functions through ERK1/2 signaling. Bone. 2021;146: 115875.PubMedCrossRef Wang B, Jiang B, Li Y, Dai Y, Li P, Li L, et al. AKAP2 overexpression modulates growth plate chondrocyte functions through ERK1/2 signaling. Bone. 2021;146: 115875.PubMedCrossRef
73.
Zurück zum Zitat Jiang HC, Chen XR, Sun HF, Nie YW. Tumor promoting effects of glucagon receptor: a promising biomarker of papillary thyroid carcinoma via regulating EMT and P38/ERK pathways. Hum Cell. 2020;33(1):175–84.PubMedCrossRef Jiang HC, Chen XR, Sun HF, Nie YW. Tumor promoting effects of glucagon receptor: a promising biomarker of papillary thyroid carcinoma via regulating EMT and P38/ERK pathways. Hum Cell. 2020;33(1):175–84.PubMedCrossRef
74.
Zurück zum Zitat Babamale AO, Chen ST. Nod-like receptors: critical intracellular sensors for host protection and cell death in microbial and parasitic infections. Int J Mol Sci. 2021;22(21):11398.PubMedPubMedCentralCrossRef Babamale AO, Chen ST. Nod-like receptors: critical intracellular sensors for host protection and cell death in microbial and parasitic infections. Int J Mol Sci. 2021;22(21):11398.PubMedPubMedCentralCrossRef
75.
Zurück zum Zitat Takahashi K, Ehata S, Miyauchi K, Morishita Y, Miyazawa K, Miyazono K. Neurotensin receptor 1 signaling promotes pancreatic cancer progression. Mol Oncol. 2021;15(1):151–66.PubMedCrossRef Takahashi K, Ehata S, Miyauchi K, Morishita Y, Miyazawa K, Miyazono K. Neurotensin receptor 1 signaling promotes pancreatic cancer progression. Mol Oncol. 2021;15(1):151–66.PubMedCrossRef
76.
Zurück zum Zitat El-Saghire H, Michaux A, Thierens H, Baatout S. Low doses of ionizing radiation induce immune-stimulatory responses in isolated human primary monocytes. Int J Mol Med. 2013;32(6):1407–14.PubMedCrossRef El-Saghire H, Michaux A, Thierens H, Baatout S. Low doses of ionizing radiation induce immune-stimulatory responses in isolated human primary monocytes. Int J Mol Med. 2013;32(6):1407–14.PubMedCrossRef
77.
Zurück zum Zitat Li C, Cai J, Ge F, Wang G. TGM2 knockdown reverses cisplatin chemoresistance in osteosarcoma. Int J Mol Med. 2018;42(4):1799–808.PubMedPubMedCentral Li C, Cai J, Ge F, Wang G. TGM2 knockdown reverses cisplatin chemoresistance in osteosarcoma. Int J Mol Med. 2018;42(4):1799–808.PubMedPubMedCentral
78.
Zurück zum Zitat Chen Y, Shan Y, Lu M, Desouza N, Guo Z, Hoffman R, et al. Alox5 blockade eradicates JAK2V617F-Induced polycythemia vera in mice. Cancer Res. 2017;77(1):164–74.PubMedCrossRef Chen Y, Shan Y, Lu M, Desouza N, Guo Z, Hoffman R, et al. Alox5 blockade eradicates JAK2V617F-Induced polycythemia vera in mice. Cancer Res. 2017;77(1):164–74.PubMedCrossRef
79.
Zurück zum Zitat Yang P, Yu D, Zhou J, Zhuang S, Jiang T. TGM2 interference regulates the angiogenesis and apoptosis of colorectal cancer via Wnt/beta-catenin pathway. Cell Cycle. 2019;18(10):1122–34.PubMedPubMedCentralCrossRef Yang P, Yu D, Zhou J, Zhuang S, Jiang T. TGM2 interference regulates the angiogenesis and apoptosis of colorectal cancer via Wnt/beta-catenin pathway. Cell Cycle. 2019;18(10):1122–34.PubMedPubMedCentralCrossRef
80.
Zurück zum Zitat Zhou X, Jiang Y, Li Q, Huang Z, Yang H, Wei C. Aberrant ALOX5 activation correlates with HER2 status and mediates breast cancer biological activities through multiple mechanisms. Biomed Res Int. 2020;2020:1703531.PubMedPubMedCentralCrossRef Zhou X, Jiang Y, Li Q, Huang Z, Yang H, Wei C. Aberrant ALOX5 activation correlates with HER2 status and mediates breast cancer biological activities through multiple mechanisms. Biomed Res Int. 2020;2020:1703531.PubMedPubMedCentralCrossRef
81.
Zurück zum Zitat Xu XS, Feng ZH, Cao D, Wu H, Wang MH, Li JZ, et al. SCARF1 promotes M2 polarization of Kupffer cells via calcium-dependent PI3K-AKT-STAT3 signalling to improve liver transplantation. Cell Prolif. 2021;54(4): e13022.PubMedPubMedCentralCrossRef Xu XS, Feng ZH, Cao D, Wu H, Wang MH, Li JZ, et al. SCARF1 promotes M2 polarization of Kupffer cells via calcium-dependent PI3K-AKT-STAT3 signalling to improve liver transplantation. Cell Prolif. 2021;54(4): e13022.PubMedPubMedCentralCrossRef
82.
Zurück zum Zitat Wang F, Wang L, Qu C, Chen L, Geng Y, Cheng C, et al. Kaempferol induces ROS-dependent apoptosis in pancreatic cancer cells via TGM2-mediated Akt/mTOR signaling. BMC Cancer. 2021;21(1):396.PubMedPubMedCentralCrossRef Wang F, Wang L, Qu C, Chen L, Geng Y, Cheng C, et al. Kaempferol induces ROS-dependent apoptosis in pancreatic cancer cells via TGM2-mediated Akt/mTOR signaling. BMC Cancer. 2021;21(1):396.PubMedPubMedCentralCrossRef
83.
Zurück zum Zitat Cheng JH, Zhang WJ, Zhu JF, Cui D, Song KD, Qiang P, et al. CaMKIIgamma regulates the viability and self-renewal of acute myeloid leukaemia stem-like cells by the Alox5/NF-kappaB pathway. Int J Lab Hematol. 2021;43(4):699–706.PubMedCrossRef Cheng JH, Zhang WJ, Zhu JF, Cui D, Song KD, Qiang P, et al. CaMKIIgamma regulates the viability and self-renewal of acute myeloid leukaemia stem-like cells by the Alox5/NF-kappaB pathway. Int J Lab Hematol. 2021;43(4):699–706.PubMedCrossRef
84.
Zurück zum Zitat Novakowski KE, Huynh A, Han S, Dorrington MG, Yin C, Tu Z, et al. A naturally occurring transcript variant of MARCO reveals the SRCR domain is critical for function. Immunol Cell Biol. 2016;94(7):646–55.PubMedPubMedCentralCrossRef Novakowski KE, Huynh A, Han S, Dorrington MG, Yin C, Tu Z, et al. A naturally occurring transcript variant of MARCO reveals the SRCR domain is critical for function. Immunol Cell Biol. 2016;94(7):646–55.PubMedPubMedCentralCrossRef
85.
Zurück zum Zitat Ding X, Xu J, Wang C, Feng Q, Wang Q, Yang Y, et al. Suppression of the SAP18/HDAC1 complex by targeting TRIM56 and Nanog is essential for oncogenic viral FLICE-inhibitory protein-induced acetylation of p65/RelA, NF-kappaB activation, and promotion of cell invasion and angiogenesis. Cell Death Differ. 2019;26(10):1970–86.PubMedPubMedCentralCrossRef Ding X, Xu J, Wang C, Feng Q, Wang Q, Yang Y, et al. Suppression of the SAP18/HDAC1 complex by targeting TRIM56 and Nanog is essential for oncogenic viral FLICE-inhibitory protein-induced acetylation of p65/RelA, NF-kappaB activation, and promotion of cell invasion and angiogenesis. Cell Death Differ. 2019;26(10):1970–86.PubMedPubMedCentralCrossRef
86.
87.
Zurück zum Zitat Wang XF, En-Zhou J, Li DJ, Mao CY, He Q, Zhang JF, et al. VSTM1 regulates monocyte/macrophage function via the NF-kappaB signaling pathway. Open Med (Wars). 2021;16(1):1513–24.CrossRef Wang XF, En-Zhou J, Li DJ, Mao CY, He Q, Zhang JF, et al. VSTM1 regulates monocyte/macrophage function via the NF-kappaB signaling pathway. Open Med (Wars). 2021;16(1):1513–24.CrossRef
88.
Zurück zum Zitat Liu Y, Duan Y, Li Y. Integrated gene expression profiling analysis reveals probable molecular mechanism and candidate biomarker in Anti-TNFalpha non-response IBD patients. J Inflamm Res. 2020;13:81–95.PubMedPubMedCentralCrossRef Liu Y, Duan Y, Li Y. Integrated gene expression profiling analysis reveals probable molecular mechanism and candidate biomarker in Anti-TNFalpha non-response IBD patients. J Inflamm Res. 2020;13:81–95.PubMedPubMedCentralCrossRef
89.
Zurück zum Zitat Wang L, Yang HY, Zang CX, Shang JM, Liu H, Zhang ZH, et al. TLR2 potentiates SR-marco-mediated neuroinflammation by interacting with the SRCR domain. Mol Neurobiol. 2021;58(11):5743–55.PubMedCrossRef Wang L, Yang HY, Zang CX, Shang JM, Liu H, Zhang ZH, et al. TLR2 potentiates SR-marco-mediated neuroinflammation by interacting with the SRCR domain. Mol Neurobiol. 2021;58(11):5743–55.PubMedCrossRef
90.
Zurück zum Zitat Wu X, Xu L, Shen Y, Yu N, Zhang Y, Guo T. MALP-2, an agonist of TLR6, promotes the immune status without affecting the differentiation capacity of umbilical cord mesenchymal stem cells. Exp Ther Med. 2017;14(6):5540–6.PubMedPubMedCentral Wu X, Xu L, Shen Y, Yu N, Zhang Y, Guo T. MALP-2, an agonist of TLR6, promotes the immune status without affecting the differentiation capacity of umbilical cord mesenchymal stem cells. Exp Ther Med. 2017;14(6):5540–6.PubMedPubMedCentral
91.
Zurück zum Zitat Matsumoto K, Yasuoka H, Yoshimoto K, Suzuki K, Takeuchi T. Platelet CXCL4 mediates neutrophil extracellular traps formation in ANCA-associated vasculitis. Sci Rep. 2021;11(1):222.PubMedPubMedCentralCrossRef Matsumoto K, Yasuoka H, Yoshimoto K, Suzuki K, Takeuchi T. Platelet CXCL4 mediates neutrophil extracellular traps formation in ANCA-associated vasculitis. Sci Rep. 2021;11(1):222.PubMedPubMedCentralCrossRef
92.
Zurück zum Zitat Wang Y, Skibbe JR, Hu C, Dong L, Ferchen K, Su R, et al. ALOX5 exhibits anti-tumor and drug-sensitizing effects in MLL-rearranged leukemia. Sci Rep. 2017;7(1):1853.PubMedPubMedCentralCrossRef Wang Y, Skibbe JR, Hu C, Dong L, Ferchen K, Su R, et al. ALOX5 exhibits anti-tumor and drug-sensitizing effects in MLL-rearranged leukemia. Sci Rep. 2017;7(1):1853.PubMedPubMedCentralCrossRef
93.
Zurück zum Zitat Sadeghi B, Al-Chaqmaqchi H, Al-Hashmi S, Brodin D, Hassan Z, Abedi-Valugerdi M, et al. Early-phase GVHD gene expression profile in target versus non-target tissues: kidney, a possible target? Bone Marrow Transplant. 2013;48(2):284–93.PubMedCrossRef Sadeghi B, Al-Chaqmaqchi H, Al-Hashmi S, Brodin D, Hassan Z, Abedi-Valugerdi M, et al. Early-phase GVHD gene expression profile in target versus non-target tissues: kidney, a possible target? Bone Marrow Transplant. 2013;48(2):284–93.PubMedCrossRef
94.
Zurück zum Zitat Normand S, Waldschmitt N, Neerincx A, Martinez-Torres RJ, Chauvin C, Couturier-Maillard A, et al. Proteasomal degradation of NOD2 by NLRP12 in monocytes promotes bacterial tolerance and colonization by enteropathogens. Nat Commun. 2018;9(1):5338.PubMedPubMedCentralCrossRef Normand S, Waldschmitt N, Neerincx A, Martinez-Torres RJ, Chauvin C, Couturier-Maillard A, et al. Proteasomal degradation of NOD2 by NLRP12 in monocytes promotes bacterial tolerance and colonization by enteropathogens. Nat Commun. 2018;9(1):5338.PubMedPubMedCentralCrossRef
95.
Zurück zum Zitat Woods B, Chen W, Chiu S, Marinaccio C, Fu C, Gu L, et al. Activation of JAK/STAT signaling in megakaryocytes sustains myeloproliferation in vivo. Clin Cancer Res. 2019;25(19):5901–12.PubMedPubMedCentralCrossRef Woods B, Chen W, Chiu S, Marinaccio C, Fu C, Gu L, et al. Activation of JAK/STAT signaling in megakaryocytes sustains myeloproliferation in vivo. Clin Cancer Res. 2019;25(19):5901–12.PubMedPubMedCentralCrossRef
96.
Zurück zum Zitat Wang Y, Zheng N, Sun T, Zhao H, Chen Y, Liu C. Role of TGM2 in Tcell lymphoblastic lymphoma via regulation of IL6/JAK/STAT3 signalling. Mol Med Rep. 2022;25(3):1–9.CrossRef Wang Y, Zheng N, Sun T, Zhao H, Chen Y, Liu C. Role of TGM2 in Tcell lymphoblastic lymphoma via regulation of IL6/JAK/STAT3 signalling. Mol Med Rep. 2022;25(3):1–9.CrossRef
97.
Zurück zum Zitat Weigert A, Strack E, Snodgrass RG, Brune B. MPGES-1 and ALOX5/-15 in tumor-associated macrophages. Cancer Metastasis Rev. 2018;37(2–3):317–34.PubMedCrossRef Weigert A, Strack E, Snodgrass RG, Brune B. MPGES-1 and ALOX5/-15 in tumor-associated macrophages. Cancer Metastasis Rev. 2018;37(2–3):317–34.PubMedCrossRef
98.
Zurück zum Zitat Nimmerjahn F, Ravetch JV. Fcγ receptors as regulators of immune responses. Nat Rev Immunol. 2008;8(1):34–47.PubMedCrossRef Nimmerjahn F, Ravetch JV. Fcγ receptors as regulators of immune responses. Nat Rev Immunol. 2008;8(1):34–47.PubMedCrossRef
99.
Zurück zum Zitat La Fleur L, Botling J, He F, Pelicano C, Zhou C, He C, et al. Targeting MARCO and IL37R on immunosuppressive macrophages in lung cancer blocks regulatory t cells and supports cytotoxic lymphocyte function. Cancer Res. 2021;81(4):956–67.PubMedCrossRef La Fleur L, Botling J, He F, Pelicano C, Zhou C, He C, et al. Targeting MARCO and IL37R on immunosuppressive macrophages in lung cancer blocks regulatory t cells and supports cytotoxic lymphocyte function. Cancer Res. 2021;81(4):956–67.PubMedCrossRef
100.
Zurück zum Zitat Gharagozloo M, Mahmoud S, Simard C, Mahvelati TM, Amrani A, Gris D. The dual immunoregulatory function of Nlrp12 in T cell-mediated immune response: lessons from experimental autoimmune encephalomyelitis. Cells. 2018;7(9):119.PubMedCentralCrossRef Gharagozloo M, Mahmoud S, Simard C, Mahvelati TM, Amrani A, Gris D. The dual immunoregulatory function of Nlrp12 in T cell-mediated immune response: lessons from experimental autoimmune encephalomyelitis. Cells. 2018;7(9):119.PubMedCentralCrossRef
101.
Zurück zum Zitat Ramirez-Ortiz ZG, Pendergraft WR, Prasad A, Byrne MH, Iram T, Blanchette CJ, et al. The scavenger receptor SCARF1 mediates the clearance of apoptotic cells and prevents autoimmunity. Nat Immunol. 2013;14(9):917–26.PubMedPubMedCentralCrossRef Ramirez-Ortiz ZG, Pendergraft WR, Prasad A, Byrne MH, Iram T, Blanchette CJ, et al. The scavenger receptor SCARF1 mediates the clearance of apoptotic cells and prevents autoimmunity. Nat Immunol. 2013;14(9):917–26.PubMedPubMedCentralCrossRef
102.
Zurück zum Zitat Mitra N, Banda K, Altheide TK, Schaffer L, Johnson-Pais TL, Beuten J, et al. SIGLEC12, a human-specific segregating (pseudo)gene, encodes a signaling molecule expressed in prostate carcinomas. J Biol Chem. 2011;286(26):23003–11.PubMedPubMedCentralCrossRef Mitra N, Banda K, Altheide TK, Schaffer L, Johnson-Pais TL, Beuten J, et al. SIGLEC12, a human-specific segregating (pseudo)gene, encodes a signaling molecule expressed in prostate carcinomas. J Biol Chem. 2011;286(26):23003–11.PubMedPubMedCentralCrossRef
103.
Zurück zum Zitat Lai TS, Greenberg CS. Histaminylation of fibrinogen by tissue transglutaminase-2 (TGM-2): potential role in modulating inflammation. Amino Acids. 2013;45(4):857–64.PubMedCrossRef Lai TS, Greenberg CS. Histaminylation of fibrinogen by tissue transglutaminase-2 (TGM-2): potential role in modulating inflammation. Amino Acids. 2013;45(4):857–64.PubMedCrossRef
104.
Zurück zum Zitat Qu J, Jiang M, Wang L, Zhao D, Qin K, Wang Y, et al. Mechanism and potential predictive biomarkers of immune checkpoint inhibitors in NSCLC. Biomed Pharmacother. 2020;127: 109996.PubMedCrossRef Qu J, Jiang M, Wang L, Zhao D, Qin K, Wang Y, et al. Mechanism and potential predictive biomarkers of immune checkpoint inhibitors in NSCLC. Biomed Pharmacother. 2020;127: 109996.PubMedCrossRef
105.
Zurück zum Zitat Basu B, Dean E, Puglisi M, Greystoke A, Ong M, Burke W, et al. First-in-human pharmacokinetic and pharmacodynamic study of the dual m-TORC 1/2 inhibitor AZD2014. Clin Cancer Res. 2015;21(15):3412–9.PubMedPubMedCentralCrossRef Basu B, Dean E, Puglisi M, Greystoke A, Ong M, Burke W, et al. First-in-human pharmacokinetic and pharmacodynamic study of the dual m-TORC 1/2 inhibitor AZD2014. Clin Cancer Res. 2015;21(15):3412–9.PubMedPubMedCentralCrossRef
106.
Zurück zum Zitat Powell JD, Pollizzi KN, Heikamp EB, Horton MR. Regulation of immune responses by mTOR. Annu Rev Immunol. 2012;30(1):39–68.PubMedCrossRef Powell JD, Pollizzi KN, Heikamp EB, Horton MR. Regulation of immune responses by mTOR. Annu Rev Immunol. 2012;30(1):39–68.PubMedCrossRef
107.
Zurück zum Zitat Albert MH, Mannert J, Fleischmann KK, Schiemann M, Pagel P, Schmid I, et al. MiRNome and transcriptome aided pathway analysis in human regulatory T cells. Genes Immun. 2014;15(5):303–12.PubMedCrossRef Albert MH, Mannert J, Fleischmann KK, Schiemann M, Pagel P, Schmid I, et al. MiRNome and transcriptome aided pathway analysis in human regulatory T cells. Genes Immun. 2014;15(5):303–12.PubMedCrossRef
108.
Zurück zum Zitat Guichard SM, Curwen J, Bihani T, D’Cruz CM, Yates JW, Grondine M, et al. AZD2014, an inhibitor of mTORC1 and mTORC2, is highly effective in ER+ breast cancer when administered using intermittent or continuous schedules. Mol Cancer Ther. 2015;14(11):2508–18.PubMedCrossRef Guichard SM, Curwen J, Bihani T, D’Cruz CM, Yates JW, Grondine M, et al. AZD2014, an inhibitor of mTORC1 and mTORC2, is highly effective in ER+ breast cancer when administered using intermittent or continuous schedules. Mol Cancer Ther. 2015;14(11):2508–18.PubMedCrossRef
109.
Zurück zum Zitat Liao H, Huang Y, Guo B, Liang B, Liu X, Ou H, et al. Dramatic antitumor effects of the dual mTORC1 and mTORC2 inhibitor AZD2014 in hepatocellular carcinoma. Am J Cancer Res. 2015;5(1):125–39.PubMed Liao H, Huang Y, Guo B, Liang B, Liu X, Ou H, et al. Dramatic antitumor effects of the dual mTORC1 and mTORC2 inhibitor AZD2014 in hepatocellular carcinoma. Am J Cancer Res. 2015;5(1):125–39.PubMed
110.
Zurück zum Zitat Axten JM, Medina JR, Feng Y, Shu A, Romeril SP, Grant SW, et al. Discovery of 7-methyl-5-(1-{[3-(trifluoromethyl)phenyl]acetyl}-2,3-dihydro-1H -indol-5-yl)-7H -pyrrolo[2,3-d ]pyrimidin-4-amine (GSK2606414), a potent and selective first-in-class inhibitor of protein kinase R (PKR)-like endoplasmic reticulum kinase (PERK). J Med Chem. 2012;55(16):7193–207.PubMedCrossRef Axten JM, Medina JR, Feng Y, Shu A, Romeril SP, Grant SW, et al. Discovery of 7-methyl-5-(1-{[3-(trifluoromethyl)phenyl]acetyl}-2,3-dihydro-1H -indol-5-yl)-7H -pyrrolo[2,3-d ]pyrimidin-4-amine (GSK2606414), a potent and selective first-in-class inhibitor of protein kinase R (PKR)-like endoplasmic reticulum kinase (PERK). J Med Chem. 2012;55(16):7193–207.PubMedCrossRef
111.
Zurück zum Zitat Rozpedek W, Pytel D, Mucha B, Leszczynska H, Diehl JA, Majsterek I. The role of the PERK/eIF2α/ATF4/CHOP signaling pathway in tumor progression during endoplasmic reticulum stress. Curr Mol Med. 2016;16(6):533–44.PubMedPubMedCentralCrossRef Rozpedek W, Pytel D, Mucha B, Leszczynska H, Diehl JA, Majsterek I. The role of the PERK/eIF2α/ATF4/CHOP signaling pathway in tumor progression during endoplasmic reticulum stress. Curr Mol Med. 2016;16(6):533–44.PubMedPubMedCentralCrossRef
112.
Zurück zum Zitat Li J, Xu J, Li Z. Obatoclax, the pan-Bcl-2 inhibitor sensitizes hepatocellular carcinoma cells to promote the anti-tumor efficacy in combination with immune checkpoint blockade. Transl Oncol. 2021;14(8):101116.PubMedPubMedCentralCrossRef Li J, Xu J, Li Z. Obatoclax, the pan-Bcl-2 inhibitor sensitizes hepatocellular carcinoma cells to promote the anti-tumor efficacy in combination with immune checkpoint blockade. Transl Oncol. 2021;14(8):101116.PubMedPubMedCentralCrossRef
113.
Zurück zum Zitat Bhati KK, Luong AM, Batoko H. VPS34 complexes in plants: untangled enough? Trends Plant Sci. 2021;26(4):303–5.PubMedCrossRef Bhati KK, Luong AM, Batoko H. VPS34 complexes in plants: untangled enough? Trends Plant Sci. 2021;26(4):303–5.PubMedCrossRef
114.
Zurück zum Zitat Noman MZ, Parpal S, Van Moer K, Xiao M, Yu Y, Arakelian T, et al. Inhibition of Vps34 reprograms cold into hot inflamed tumors and improves anti–PD-1/PD-L1 immunotherapy. Sci Adv. 2020;6(18):x7881.CrossRef Noman MZ, Parpal S, Van Moer K, Xiao M, Yu Y, Arakelian T, et al. Inhibition of Vps34 reprograms cold into hot inflamed tumors and improves anti–PD-1/PD-L1 immunotherapy. Sci Adv. 2020;6(18):x7881.CrossRef
115.
Zurück zum Zitat Jiang J, Jin Z, Zhang Y, Peng L, Zhang Y, Zhu Z, et al. Robust prediction of immune checkpoint inhibition therapy for non-small cell lung cancer. Front Immunol. 2021;12:646874.PubMedPubMedCentralCrossRef Jiang J, Jin Z, Zhang Y, Peng L, Zhang Y, Zhu Z, et al. Robust prediction of immune checkpoint inhibition therapy for non-small cell lung cancer. Front Immunol. 2021;12:646874.PubMedPubMedCentralCrossRef
Metadaten
Titel
Construction of a predictive model for immunotherapy efficacy in lung squamous cell carcinoma based on the degree of tumor-infiltrating immune cells and molecular typing
verfasst von
Lingge Yang
Shuli Wei
Jingnan Zhang
Qiongjie Hu
Wansong Hu
Mengqing Cao
Long Zhang
Yongfang Wang
Pingli Wang
Kai Wang
Publikationsdatum
01.12.2022
Verlag
BioMed Central
Erschienen in
Journal of Translational Medicine / Ausgabe 1/2022
Elektronische ISSN: 1479-5876
DOI
https://doi.org/10.1186/s12967-022-03565-7

Weitere Artikel der Ausgabe 1/2022

Journal of Translational Medicine 1/2022 Zur Ausgabe

Leitlinien kompakt für die Innere Medizin

Mit medbee Pocketcards sicher entscheiden.

Seit 2022 gehört die medbee GmbH zum Springer Medizin Verlag

Echinokokkose medikamentös behandeln oder operieren?

06.05.2024 DCK 2024 Kongressbericht

Die Therapie von Echinokokkosen sollte immer in spezialisierten Zentren erfolgen. Eine symptomlose Echinokokkose kann – egal ob von Hunde- oder Fuchsbandwurm ausgelöst – konservativ erfolgen. Wenn eine Op. nötig ist, kann es sinnvoll sein, vorher Zysten zu leeren und zu desinfizieren. 

Umsetzung der POMGAT-Leitlinie läuft

03.05.2024 DCK 2024 Kongressbericht

Seit November 2023 gibt es evidenzbasierte Empfehlungen zum perioperativen Management bei gastrointestinalen Tumoren (POMGAT) auf S3-Niveau. Vieles wird schon entsprechend der Empfehlungen durchgeführt. Wo es im Alltag noch hapert, zeigt eine Umfrage in einem Klinikverbund.

Proximale Humerusfraktur: Auch 100-Jährige operieren?

01.05.2024 DCK 2024 Kongressbericht

Mit dem demographischen Wandel versorgt auch die Chirurgie immer mehr betagte Menschen. Von Entwicklungen wie Fast-Track können auch ältere Menschen profitieren und bei proximaler Humerusfraktur können selbst manche 100-Jährige noch sicher operiert werden.

Die „Zehn Gebote“ des Endokarditis-Managements

30.04.2024 Endokarditis Leitlinie kompakt

Worauf kommt es beim Management von Personen mit infektiöser Endokarditis an? Eine Kardiologin und ein Kardiologe fassen die zehn wichtigsten Punkte der neuen ESC-Leitlinie zusammen.

Update Innere Medizin

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.