Abstract
Background
Skin cutaneous malignant melanoma (SKCM) is among the most aggressive
forms of skin cancer, notorious for its rapid progression and poor
prognosis under late diagnosis. This study investigates the role of
circadian rhythm-related genes (CRGs) in SKCM addressing a gap in
understanding how CRGs affect tumor progression and patient outcomes.
Methods
An analysis of CRGs expression was conducted on SKCM samples derived
from The Cancer Genome Atlas datasets(TCGA). Moreover, a correlation
between various subtypes and their clinical features was identified.
The study employed various bioinformatics methods, including
differential expression analysis, consensus clustering, and survival
analysis, to investigate the role of CRGs. The functional consequences
of various CRG expression patterns were further investigated using
immune infiltration analysis and gene set variation analysis (GSVA). A
scoring system based on CRGs was developed to predict overall survival
(OS) and treatment responses in SKCM patients. The predictive accuracy
of this score system was then tested, and a nomogram was used to
improve its clinical usefulness.
Results
Key findings from this study include significant genetic alterations in
circadian rhythm-related genes (CRGs) in skin cutaneous melanoma
(SKCM), such as mutations and CNVs. Two molecular subtypes with
distinct clinical outcomes and immune profiles were identified. A
prognostic model based on six CRGs (CMTM, TNPO1, CTBS, UTRN, HK2, and
LIF) was developed and validated with TCGA and GEO datasets, showing
high predictive accuracy for overall survival (OS). A high CRGs score
correlated with poor OS, immune checkpoint expression, and reduced
sensitivity to several chemotherapeutics, including AKT inhibitor VIII
and Camptothecin.
Conclusions
This work provides valuable insights into the circadian regulation of
SKCM and underscores the potential of CRGs as biomarkers for prognosis
and targets for therapeutic interventions. The novel molecular subtypes
and CRGs prognostic scoring model introduced in this study offer
significant contributions to the understanding and management of SKCM.
Keywords: melanoma, circadian rhythm related genes, prognosis, tumor
immune microenvironment, immunotherapy
1. Introduction
Cutaneous malignant melanoma (SKCM) is a highly aggressive skin cancer
originating from melanocytes. In recent years, the incidence of SKCM
has been on the rise, particularly in Western countries, where it has
become one of the fastest-growing malignancies. Recent cancer
statistics indicate that in 2023, the United States is projected to see
approximately 97,610 new melanoma cases and 7,990 associated fatalities
([33]1). Projections suggest a substantial increase, with potential
rises to 510,000 new cases and 96,000 deaths by 2040 if current trends
persist ([34]2). Although early detection and treatment could
significantly improve survival rates, the prognosis remains poor once
metastasis occurs, with a low five-year survival rate ([35]3).
Immunotherapy has emerged as a key therapeutic option for metastatic
melanoma since the development of immune checkpoint inhibitors, such as
anti-PD-1 and anti-CTLA-4 antibodies ([36]4). However, despite the
significant therapeutic effects observed in some patients, a
considerable proportion does not respond to treatment or eventually
develop resistance, greatly limiting the widespread clinical
application of immunotherapy ([37]5). Therefore, identifying reliable
biomarkers to predict patient responses to immunotherapy and to develop
personalized treatment strategies is of paramount importance.
Circadian rhythm genes (CRGs) regulate various physiological processes
by controlling the cell cycle, DNA repair, and metabolism ([38]6). In
recent years, increasing research has revealed the crucial role of CRGs
in cancer development, progression, and treatment response ([39]7).
Dysregulated CRG expression is strongly linked to tumor growth and a
poor prognosis in many forms of cancer. For example, studies have shown
that changes in CRG expression in breast cancer could affect cell
proliferation and apoptosis, thereby influencing patient outcomes
([40]8). Although research on CRGs in melanoma is relatively limited,
existing evidence suggests that these genes might play a role in tumor
development by regulating the tumor microenvironment, immune evasion,
and metabolic reprogramming ([41]9). Thus, exploring the mechanisms by
which CRGs function in melanoma, particularly in the context of
immunotherapy, holds significant theoretical and clinical implications.
Recent studies have increasingly illuminated the crucial role of CRGs
in the field of cancer immunotherapy, emphasizing their potential to
revolutionize treatment modalities for diseases like melanoma, which is
heavily dependent on immunological strategies for management and cure
([42]10). Specifically, CRGs might influence tumor responses to
immunotherapy by regulating the circadian rhythms of the immune system.
For instance, key CRGs such as CLOCK and BMAL1 have been shown to
modulate the functional state of immune cells within the
tumor microenvironment, thereby affecting the efficacy of immunotherapy
([43]11). This modulation is crucial because it can considerably
improve immunotherapy efficacy by matching treatment time to the
patient’s biological clocks, possibly enhancing therapeutic
effectiveness while reducing side effects. Moreover, the expression
patterns of CRGs might be closely related to patient prognosis,
offering possibilities for the development of prognostic models based
on CRGs. In melanoma, given its strong reliance on immunotherapy,
focusing on how CRGs affect melanoma’s response to immunotherapy, this
research could pave the way for novel, more effective treatment
protocols that are synchronized with the patient’s circadian biology,
offering a new dimension to oncological care and a promising avenue for
future scientific exploration and clinical application.
Building on this background, this study aims to systematically analyze
the expression characteristics of CRGs and their relationships with
immunotherapy response and patient prognosis by integrating clinical
and gene expression data from a large cohort of SKCM patients. We
hypothesize that CRGs expression patterns predict patient responses to
immunotherapy, and serve as independent prognostic markers, providing
new foundations for the precision treatment of melanoma. The results of
this study are expected to support personalized treatment strategies
for melanoma and expand the application prospects of CRGs in cancer
immunotherapy.
2. Materials and methods
2.1. Data acquisition and circadian rhythm-related genes
The Cancer Genome Atlas (TCGA) ([44]https://portal.gdc.cancer.gov/),
the Genotype-Tissue Expression Project (GTEx)
([45]https://www.genome.gov/Funded-Programs-Projects/Genotype-Tissue-Ex
pression-Project), and the Gene Expression Omnibus (GEO)
([46]https://www.ncbi.nlm.nih.gov/geo) databases provided data on gene
expression and clinical pathology for Skin Cutaneous Melanoma (SKCM).
All samples in the datasets were derived from Homo sapiens, with
patients lacking survival information excluded. To enhance the
consistency and reliability of the data, we integrated gene expression
data from the TCGA-SKCM (The Cancer Genome Atlas - Skin Cutaneous
Melanoma) and [47]GSE65904 datasets. Specifically, the TCGA provided
FPKM (Fragments Per Kilobase of transcript per Million mapped reads)
sequencing data for 471 SKCM patients, combined with corresponding
clinical information, which served as an essential foundation for our
subsequent analyses ([48]12) ([49] Table 1 ), and transcriptome data
for 812 normal skin samples from the GTEx database ([50]13). The
[51]GSE65904 dataset, based on the [52]GPL10558 platform, comprises
transcriptome data from 214 SKCM patients of Homo sapiens origin
([53]14, [54]15). During the processing of dataset [55]GSE65904,
normalization was performed by applying a log transformation to the
expression data to eliminate skewness and enhance comparability.
Additionally, the avereps function was used to average genes with
repeated measurements, thereby reducing technical noise.
Standardization was carried out using the normalizeBetweenArrays
function from the limma package ([56]16) to adjust for technical
variations between samples, ensuring data consistency and
comparability. These steps ensured the reliability and accuracy of the
subsequent analysis results. During the data integration process,
differences in experimental conditions, technical platforms, and sample
handling methods between datasets may lead to expression biases, known
as batch effects. These inconsistencies could negatively affect the
subsequent analysis results. Therefore, we applied the “Combat”
algorithm to correct for batch effects in the data. The Combat
algorithm uses parametric modeling to evaluate and adjust the impact of
batch effects on gene expression data, thereby eliminating systematic
biases between batches. This correction ensured that data from
different sources could be compared under the same standard, providing
a more robust foundation for downstream analyses. After correcting for
batch effects, we generated a merged transcriptomic dataset that
effectively eliminated potential inconsistencies caused by data
discrepancies, ensuring the accuracy and reliability of statistical
analyses. This process is crucial for subsequent analyses, including
differential expression analysis, gene correlation studies, and
prognostic evaluations. The TCGA-SKCM dataset’s FPKM values were
translated to transcripts per million (TPM). After merging the
TCGA-SKCM and [57]GSE65904 datasets, the “Combat” method was used to
eliminate batch effects ([58] Supplementary Figure S1 ).
Table 1.
Baseline data from TCGA.
Characteristic levels TCGA-SKCM
n 471
Gender, n (%) Female 179 (38%)
Male 292 (62%)
Age, n (%) <=60 252 (54.4%)
>60 211 (45.6%)
T stage, n (%) T1 41 (11.3%)
T2 79 (21.7%)
T3 91 (25%)
T4 153 (42%)
N stage, n (%) N0 235 (56.8%)
N1 74 (17.9%)
N2 49 (11.8%)
N3 56 (13.5%)
M stage, n (%) M0 418 (94.4%)
M1 25 (5.6%)
Pathologic stage, n (%) Stage I 77 (18.7%)
Stage II 140 (34%)
Stage III 171 (41.5%)
Stage IV 24 (5.8%)
OS event, n (%) Alive 247 (53.2%)
Dead 217 (46.8%)
DSS event, n (%) Alive 267 (58.3%)
Dead 191 (41.7%)
PFI event, n (%) Alive 153 (33%)
Dead 311 (67%)
[59]Open in a new tab
To investigate somatic mutations, 469 SKCM patients’ “Masked Somatic
Mutation” data were chosen from the TCGA GDC website. VarScan software
was used to preprocess the data, and the maftools program was used to
show somatic mutations ([60]17). We downloaded the “Masked Copy Number
Segment” data for 471 SKCM patients in order to examine copy number
variations (CNVs) in important genes among TCGA-SKCM patients. GISTIC
2.0 was used to evaluate the CNV segment data ([61]18). In GenePattern
([62]https://cloud.genepattern.org) to investigate CNVs in circadian
rhythm-related genes (CRGs).
Circadian rhythm-related key genes (CRGs) were obtained from previous
studies and reviews ([63]19, [64]20). Aiming to represent the core
regulatory roles of circadian rhythms in various biological processes,
we extracted 17 CRGs for further research, including ARNTL, ARNTL2,
CLOCK, CRY1, CRY2, CSNK1D, CSNK1E, NPAS2, NR1D1, NR1D2, PER1, PER2,
PER3, RORA, RORB, RORC, and TIMELESS. In this study, the selection of
17 CRGs was based on their well-established roles in circadian
regulation and cancer research. Circadian rhythm genes are highly
conserved throughout evolution, from invertebrates to humans, and play
critical roles in regulating physiological functions and
disease-related pathways. The same foundational principle applies to
melanoma, as circadian rhythm disruptions can influence cancer cell
cycles and growth ([65]19). Therefore, the selection of these genes
primarily considered their functional conservation in model organisms.
Additionally, the impact of these genes in prostate cancer,
particularly their key roles in the tumor immune microenvironment, was
also taken into account ([66]20). Our selection process incorporated
similar bioinformatics analyses, including the use of TCGA and
GeneCards databases, along with Lasso and Cox regression analyses, to
confirm the expression patterns of these genes in melanoma samples and
their alignment with circadian rhythm regulation characteristics.
Special attention was given to genes that have demonstrated significant
influence on tumor progression and immune regulation in multiple
studies. As a result, the selected CRGs not only represent key
molecules in the circadian regulation of melanoma but may also
influence melanoma growth and development by modulating the tumor
immune microenvironment. The selection of these genes was guided by
their demonstrated significant biological functions in prior research,
with the aim of further uncovering the roles of circadian rhythms in
tumor progression.
2.2. Differential expression, gene correlation, and prognosis analysis
To further investigate the expression characteristics of circadian
rhythm-related genes (CRGs) in SKCM samples, we integrated the
TCGA-SKCM and GTEx datasets and performed differential expression
analysis between the tumor group (SKCM) and the normal group (normal
skin tissues) using the limma package. During the data analysis, we
evaluated the expression levels of each gene and identified those
significantly upregulated or downregulated in tumor tissues. The
results of the differential expression analysis were presented as
boxplots, which visually depict the significant differences in gene
expression between the two groups. Subsequently, to explore the
relationships among CRGs, we performed Spearman rank correlation
analysis. This method calculates correlation coefficients between gene
expression levels, revealing significant correlations and regulatory
networks among the genes. Using the ggplot2 package for visualization,
we generated a correlation heatmap that clearly illustrates the
expression patterns among different CRGs, providing valuable insights
into their potential biological functions. Based on the expression
levels of CRGs in SKCM, we further divided the samples into a
low-expression group (0%-50%) and a high-expression group (51%-100%).
This grouping strategy facilitates the analysis of how different
expression levels of CRGs impact patient prognosis. Subsequently, we
used the survminer and survival packages to analyze the prognostic
differences between these two groups. We applied a Cox regression model
to evaluate survival risks between the low-expression and
high-expression groups and used the log-rank test to compare the
survival curves between groups.
2.3. Identification of circadian rhythm-related molecular subtypes
One technique for figuring out the size and composition of possible
clusters in a dataset (gene expression profiles) is consensus
clustering ([67]21). In order to distinguish between distinct circadian
rhythm-associated subtypes, we used the “ConsensusClusterPlus” package
to conduct consensus clustering on the combined TCGA and GEO datasets
utilizing important genes linked to circadian rhythm ([68]21).
ClusterAlg = “pam” and distance = “euclidean” were used to sample 80%
of the total data in 100 repetitions, with the number of clusters being
set between 2 and 9.
2.4. Relationship between circadian rhythm-related molecular subtypes and
clinical features
We examined the connections between molecular subtypes, clinical
pathological features, gene expression levels, molecular functions, and
immunological infiltration in order to investigate the clinical use of
the two subtypes found by consensus clustering. Patient characteristics
included age, gender, and TNM staging. Additionally, we downloaded the
“c5.go.bp.v7.5.1.symbols” and “c2.cp.kegg.v7.4.symbols” gene sets from
the MSigDB database ([69]22) to perform gene set variation analysis
(GSVA). Gene expression matrices from various samples were converted
into gene set expression matrices using GSVA, a non-parametric and
unsupervised analytic technique, in order to assess gene set enrichment
findings from transcriptome microarrays ([70]23). Based on the gene
expression matrix of each sample, pathway scores were computed using
the GSVA package in R ([71]https://github.com/rcastelo/GSVA), and the
limma program was used to perform differential screening of enriched
functions or pathways.
Additionally, we measured the relative abundance of immune cell
infiltration for every sample using the ssGSEA method ([72]24). There
are 28 different kinds of immune cells that have been discovered,
including natural killer T cells, macrophages, activated CD8 T cells,
and activated B cells. The relative abundance of each kind of immune
cell in the samples was represented by enrichment scores that were
computed using ssGSEA. The ggplot2 software was used to visualize the
association between immune cell expression and several molecular
subtypes.
2.5. Identification of differentially expressed genes (DEGs) between
molecular subtypes
Using the R limma package, differential analysis of genes between
several molecular subtypes was carried out. The criterion of a fold
change more than 1.5 and an adjusted p-value less than 0.05 was used to
identify differentially expressed genes (DEGs) among subtypes of
circadian rhythms. In order to maximize the inclusion of DEGs with
practical biological significance in terms of expression levels, we
selected |logFC| > 0.585 and adjusted P value < 0.05 as the screening
criteria ([73]25). DEGs that were found to be up-regulated were those
with logFC > 0.585 and adj p value < 0.05, whereas down-regulated DEGs
were identified as those with logFC < -0.585 and adj p value < 0.05.
2.6. Functional enrichment analysis of differentially expressed genes between
circadian rhythm subtypes
A popular technique for conducting extensive functional enrichment
studies on biological processes (BP), molecular functions (MF), and
cellular components (CC) is gene ontology (GO) analysis ([74]26). The
Kyoto Encyclopedia of Genes and Genomes (KEGG) is an extensive database
that stores data about illnesses, medications, biological processes,
and genomes ([75]27). Using the clusterProfiler package ([76]28), GO
annotation analysis and KEGG pathway enrichment analysis of DEGs were
carried out ([77]27), with FDR < 0.05 being regarded as statistically
significant. Adj p.value < 0.05 and q.value < 0.05 were the selection
criterion for entries, and the Benjamini-Hochberg technique (BH) was
used to correct the p-value.
2.7. Identification of circadian rhythm-related gene clusters and
construction of a prognostic model
To further search prognostic genes and construct a prognosis model,
multivariate Cox regression analysis, lasso regression analysis, and
univariate Cox regression analysis were used. Initially, univariate Cox
regression analysis was carried out on DEGs, keeping genes linked with
SKCM prognosis if p < 0.05. Consensus clustering was used to classify
SKCM patients into distinct gene clusters, and then these gene clusters
were examined for variations in gene expression and prognosis. Based on
the combined TCGA and GEO datasets, all SKCM patients were randomly
divided into training and test sets in a 1:1 ratio. The LASSO technique
([78]29) was utilized in the training set to reduce multicollinearity
and filter significant variables from univariate Cox regression
analysis. To acquire more accurate independent prognostic indicators
(prognostic characteristic genes), we applied multivariate Cox
regression analysis with stepwise regression for final screening.
Lastly, by considering the optimized gene expression and corresponding
estimated Cox regression coefficients, we generated a risk score
formula:
[MATH:
Risksco
re=(exp−Gene1×coef−Gene1)+(exp−Gene2×coef−Gene2)+…+(exp−Gene×coef−Gene<
mo stretchy="false">) :MATH]
. The training and test set samples were separated into high-risk and
low-risk categories based on their median risk score ([79]30). The term
exp-Gene represents the expression level of a specific gene, and
coef-Gene represents the regression coefficient of that gene. To
compare overall survival between the training and test sets,
Kaplan-Meier analysis was performed using the survival program.
Furthermore, survival prediction was assessed using time-dependent
receiver operating characteristic (ROC) curves, and prognostic or
predictive accuracy was measured by calculating the area under the ROC
curve (AUC) using the timeROC R package ([80]31).
2.8. Gene Set Enrichment Analysis
We employed Gene Set Enrichment Analysis (GSEA) to perform functional
enrichment analysis in order to investigate biological process
differences between high-risk and low-risk SKCM patients. A computer
technique called GSEA assesses if a predetermined gene set exhibits
statistically significant changes between two biological states
([81]32). Usually, it is applied to quantify changes in biological
process and pathway activity within samples of expression datasets. For
Gene Set Enrichment Analysis, the gene sets “c2.cp.v7.2.symbols.gmt”
and “h.all.v7.2.symbols.gmt” were obtained from the MSigDB database
([82]22), with a false discovery rate (FDR) < 0.25 regarded as
substantially enriched.
2.9. Immune infiltration analysis
CIBERSORTx ([83]33) is a method that estimates the makeup and quantity
of immune cells inside mixed cells by deconvolution of transcriptome
expression matrices using linear support vector regression. To generate
the immune cell infiltration matrix, we uploaded gene expression matrix
data (TPM) to CIBERSORTx, used the LM22 signature gene matrix, and
filtered samples with p<0.05. The association between risk scores,
critical genes, and immune cell infiltration levels was investigated.
The ESTIMATE R package ([84]34) was utilized to predict stromal and
immune cell scores from gene expression profiles, followed by the
calculation of these cell counts. We investigated the relationship
between ESTIMATE scores and high- and low-risk categories.
2.10. Prediction of immunotherapy response and drug sensitivity
Utilizing the Tumor Immune Dysfunction and Exclusion (TIDE) tool
([85]35), prognosis was assessed based on risk scores, and variations
in TIDE, Dysfunction, and Exclusion scores between high-risk and
low-risk groups were evaluated. Additionally, immunotherapy responses
in SKCM patients were predicted. Genomic sensitivity indicators and
tumor drug response information were found using the Genomics of Drug
Sensitivity in Cancer (GDSC) database ([86]36). Using the pRRophetic
algorithm ([87]37), we constructed a ridge regression model based on
the expression profiles of CCLE cell lines and TCGA-SKCM gene
expression profiles. This allowed us to estimate the IC50 values for
the sensitivity of common anticancer treatments in both high-risk and
low-risk groups.
2.11. Construction of a nomogram scoring system
We conducted univariate and multivariate Cox regression studies on risk
scores and clinical parameters in order to optimize the prediction
power of the model and further evaluate the influence of risk scores
and clinicopathological aspects on patient prognosis. The R package
“regplot” ([88]https://CRAN.R-project.org/package=regplot) was used to
construct a nomogram by integrating clinical features and risk scores,
predicting 1-, 3-, and 5-year survival probabilities for SKCM patients.
Each variable in the nomogram scoring system has a corresponding score,
and the sum of all the variable scores for each sample yields the final
score. After that, time-dependent ROC curves and calibration curves
were used to assess the nomogram’s discriminating power.
2.12. Statistical analysis
R programming was used for all statistical analysis and data
processing. The Mann-Whitney U test (also known as the Wilcoxon
rank-sum test) was used to analyze differences between non-normally
distributed variables, while the independent Student’s t-test was used
to estimate the statistical significance of normally distributed
variables for comparisons between two groups of continuous variables. P
< 0.05 was deemed statistically significant for all two-sided
statistical p-values. A detailed flow chart of this research can be
found in [89]Figure 1 .
Figure 1.
[90]Figure 1
[91]Open in a new tab
Flowchart of the research. TCGA, The Cancer Genome Atlas; GDC, Genomic
Data Commons; SKCM, Skin cutaneous malignant melanoma; CRGs, circadian
rhythm-related genes; GTEx, Genotype-Tissue Expression Project; SNP,
Single Nucleotide Polymorphism; CNV, copy number variations; GSVA, gene
set variation analysis; ssGSEA, single sample gene set enrichment
analysis; DEGs, Differentially Expressed Genes; LASSO, Least Absolute
Shrinkage and Selection Operator; KM, Kaplan-Meier; ROC, Receiver
Operating Characteristic Curve.
3. Results
3.1. SNPs and CNVs in circadian rhythm-related genes
This study included 17 circadian rhythm-related genes (CRGs). A summary
analysis of somatic mutation rates in these 17 CRGs revealed that 123
out of 469 SKCM samples (26.23%) had mutations in CRGs. Among these,
RORB had the highest mutation frequency (6%), followed by PER3, RORC,
TIMELESS, and others ([92] Figure 2A ). Next, we investigated somatic
copy number variations (CNVs) in these CRGs ([93] Figure 2B ) and found
that CNVs were common across all 17 CRGs. Notably, CSNK1D, CSNK1E,
RORC, and CLOCK exhibited widespread increases in copy number
variations, whereas PER3, PER2, PER1, and CRY1 showed decreases in
CNVs. [94]Figure 2C illustrates the locations of CNV changes in CRGs on
their respective chromosomes, such as CNV changes in CSNK1D located on
chromosome 17 and in CSNK1E on chromosome 22.
Figure 2.
[95]Figure 2
[96]Open in a new tab
Genetic and transcriptomic alterations of CRGs. (A) Mutation
frequencies of 17 CRGs in patients from TCGA. (B) Frequency of CNV
gains and losses among CRGs. (C) Locations of CNV changes of CRGs
across 23 chromosomes. (D) Expression differences of 17 CRGs between
normal tissues and SKCM tissues. (E) PCA results of transcriptomic data
from tumor and normal tissues. *** indicates P<0.001. TCGA, The Cancer
Genome Atlas; CRGs, Circadian Rhythm Related Genes; CNV, Copy Number
Variations; SKCM, Skin Cutaneous Melanoma; PCA, Principal Component
Analysis.
3.2. Differential expression, gene correlation, and prognosis analysis
We further compared the mRNA expression levels of CRGs between tumor
tissues and normal tissues. As shown in [97]Figure 2D , compared with
normal skin tissues, the expression of ARNTL, ARNTL2, CRY1, CRY2,
CSNK1E, NR1D1, NR1D2, PER1, PER2, PER3, RORA, RORB, and RORC was
significantly downregulated in tumor tissues (P<0.001), whereas the
expression of CSNK1D and TIMELESS was significantly upregulated
(P<0.001). [98]Figure 2E shows the PCA analysis results of tumor and
normal tissues.
Next, we analyzed the expression correlation among CRGs. Spearman
correlation analysis revealed significant correlations between the
expression of the 17 genes, with most genes showing positive
correlations, except for TIMELESS and CLOCK, which were negatively
correlated with other genes ([99] Figure 3A ). [100]Figure 3B
illustrates the interactions among 17 CRGs, with each node representing
a gene and edges denoting the correlations between genes and their
associations with patient prognosis. The color and style of the edges
indicate the type of correlation: pink lines represent positive
correlations, while blue lines represent negative correlations, both
with high statistical significance (P < 0.0001). The network reveals
that most CRGs exhibit significant positive correlations, suggesting
co-expression in tumor cells and involvement in related biological
processes. In contrast, negative correlations involving TIMELESS,
CLOCK, and other genes indicate distinct functions or regulatory
mechanisms, potentially playing opposing roles in tumor progression.
This network provides insight into the complex interplay of CRGs and
their roles in melanoma biology. Moreover, [101]Figure 3C highlights
the association between gene expression and patient prognosis,
demonstrating that high expression of ARNTL2, ARNTL, NR1D2, and RORC
correlates with favorable outcomes, while elevated levels of PER1,
PER2, CRY1, TIMELESS, PER3, and CSNK1E are linked to poor prognosis. By
illustrating both the interactions among these genes and their impact
on survival time and quality of life, the figure provides valuable
insights into the prognostic significance of specific gene expressions
and the complex interactions of CRGs and their roles in melanoma
biology. Understanding the interactions between these genes and their
prognostic significance may provide new biomarkers for clinical
applications, facilitating risk stratification and personalized
treatment strategies.
Figure 3.
[102]Figure 3
[103]Open in a new tab
Gene correlation and prognosis analysis. (A) Correlation of expression
levels among 17 CRGs. (B) Prognostic network diagram. (C) Prognostic
differences between high and low expression groups of CRGs based on
TCGA data. • indicates P<0.05, ** indicates P<0.01, *** indicates
P<0.001. CRGs, Circadian Rhythm Related Genes; TCGA, The Cancer Genome
Atlas.
3.3. Identification of circadian rhythm-related molecular subtypes
In this study, to identify circadian rhythm-related molecular subtypes
of melanoma, we employed an unsupervised consensus clustering analysis.
Specifically, we used the “ConsensusClusterPlus” R package to perform
consensus clustering on the combined TCGA and GEO datasets, with 17
circadian rhythm-related genes as features. This method assesses the
stability of different clustering numbers by repeated resampling and
uses the consensus matrix to evaluate the optimal number of clusters
(k). Ultimately, we determined that k=2 was the optimal clustering
number and categorized the patients into two molecular subtypes
accordingly. [104]Figure 4A shows the consensus matrix generated
through consensus clustering analysis, where the color intensity of
each cell represents the similarity between samples. For the clustering
scenario with k=2, the samples are divided into two main groups,
illustrating the relationships between the samples and the
concentration of their clustering. [105]Figure 4B The figure displays
the consensus matrix for k=3. Compared to k=2, the sample division
becomes more complex, potentially resulting in multiple subgroups. At
this point, the similarity between samples decreases, indicating a
lower clustering quality and suggesting that the clustering may not be
optimal under these conditions. [106]Figure 4C presents the consensus
CDF curves for different k values. Compared to other k values, the
curve for k=2 has the largest proportion in the region above 0.8,
indicating it as the optimal clustering choice. This provides strong
evidence for determining the most suitable number of clusters for the
samples. [107]Figure 4D shows the number of clusters (k) on the x-axis
and the change in clustering quality (Delta area) on the y-axis. As the
k value increases, there is a clear downward trend in the Delta area,
with the most significant drop observed at k=2. This further supports
the conclusion that k=2 is the optimal choice. [108]Figures 4A–D
display the clustering results obtained at different k values based on
the consensus clustering method. After considering the consensus
cumulative distribution function (CDF) curve, Delta area variation, and
clustering stability, we ultimately selected k=2 as the optimal number
of clusters. This classification was based on the similarity of gene
expression patterns between samples, rather than traditional methods
using mean or median cutoffs, ensuring the rationality and stability of
the SKCM molecular subtypes. Significant variations in the gene
expression patterns of the two subtypes were found using PCA analysis
([109] Figure 4F ). [110]Figures 4E, G show the relationship between
CRG expression and clinicopathological features, with circadian
rhythm-related genes expressed significantly higher in CRG cluster B
than in CRG cluster A: ARNTL (P<0.001), ARNTL2 (P<0.001), CLOCK
(P<0.001), CRY1 (P<0.001), CRY2 (P<0.001), CSNK1D (P<0.001), CSNK1E
(P<0.001), NPAS2 (P<0.001), NR1D1 (P<0.001), NR1D2 (P<0.001), PER1
(P<0.001), PER2 (P<0.001), PER3 (P<0.001), RORA (P<0.001), RORB
(P<0.05), RORC (P<0.001).
Figure 4.
[111]Figure 4
[112]Open in a new tab
Identification of circadian rhythm-related molecular subtypes. (A)
Consensus clustering heatmap defining two clusters (k = 2) and their
associated area. (B) Consensus clustering heatmap defining three
clusters (k = 3) and their associated area. (C) CDF plot showing the
distribution of consensus clustering corresponding to each k (D) Delta
area plot visualizing the relative change in the area under the CDF
curve at different k values. (E) Heatmap showing the relationship
between gene expression levels and clinicopathological parameters
across different clusters. (F) PCA analysis showing differences in gene
expression profiles between the two subtypes. (G) Expression
differences of CRGs between CRG cluster A and B. • indicates P<0.05, **
indicates P<0.01, *** indicates P<0.001. CDF, Cumulative Distribution
Function; PCA, Principal Component Analysis; CRGs, Circadian Rhythm
Related Genes.
3.4. GSVA and immune infiltration analysis
GSVA enrichment analysis indicated that CRG cluster B was significantly
enriched in biological processes related to protein synthesis,
metabolism, and substance transport, including positive regulation of
response to endoplasmic reticulum stress, golgi organization, magnesium
ion transport, and er nucleus signaling pathway ([113] Figure 5A ).
Additionally, cluster B was enriched in pathways such as mtor signaling
pathway, erbb signaling pathway, insulin signaling pathway, and retinol
metabolism ([114] Figure 5B ). To further investigate the role of CRGs
in the tumor immune microenvironment (TME), we used the ssGSEA
algorithm to assess the relative abundance of immune cells in each SKCM
sample and the differences between clusters ([115] Figure 5C ). The
results showed that the immune infiltration levels of activated B
cells, activated CD8+ T cell, CD56dim natural killer cells,
Myeloid-derived suppressor cells, macrophages, monocytes, and natural
killer T cells were significantly higher in CRG cluster A than in CRG
cluster B. In contrast, immature dendritic cells, plasmacytoid
dendritic cells, and type 2 T helper cells were significantly more
infiltrated in CRG cluster B than in CRG cluster A.
Figure 5.
[116]Figure 5
[117]Open in a new tab
GSVA and immune infiltration analysis. (A) GSVA-GOBP analysis results
between CRG cluster A and B. (B) GSVA-KEGG analysis results between CRG
cluster A and B. (C) ssGSEA immune infiltration analysis showing
differences in the relative abundance of immune cells between CRG
cluster A and B. • indicates P<0.05, ** indicates P<0.01, *** indicates
P<0.001. GSVA, Gene Set Variation Analysis; GOBP, Gene Ontology
Biological Process; KEGG, Kyoto Encyclopedia of Genes and Genomes.
3.5. Identification of differentially expressed genes (DEGs) between
molecular subtypes
Using the “limma” package in R, we found 457 circadian rhythm
subtype-related DEGs and carried out functional enrichment analysis to
investigate the possible biological roles of each CRG cluster. ([118]
Figures 6A, B , [119]Table 2 ). Cell adhesion and communication-related
biological processes, including cell-matrix adhesion, cell-substrate
adhesion, cell-cell junction, integrin complex, cell adhesion molecule
binding, and extracellular matrix binding, were shown to be enriched in
DEGs, according to GO enrichment analysis. The ECM-receptor
interaction, focal adhesion, proteoglycans in cancer, PI3K-Akt
signaling pathway, and Hippo signaling pathway were among the pathways
in which DEGs were enriched, according to KEGG enrichment analysis.
Figure 6.
[120]Figure 6
[121]Open in a new tab
Identification of gene clusters based on OS-DEGs. (A) GO functional
enrichment analysis of DEGs between circadian rhythm-related molecular
subtypes. (B) KEGG functional enrichment analysis of DEGs between
circadian rhythm-related molecular subtypes. (C) Consensus clustering
heatmap defining two clusters (k = 2) and their associated area. (D)
Consensus clustering heatmap defining three clusters (k = 3) and their
associated area. (E) CDF plot showing the distribution of consensus
clustering corresponding to each k. (F) Delta area plot visualizing the
relative change in the area under the CDF curve at different k values.
(G) Heatmap showing the relationship between gene expression levels and
clinicopathological parameters across the two gene clusters. (H)
Kaplan-Meier curve showing the prognostic differences between the two
gene clusters. DEG, Differentially Expressed Gene; GO, Gene Ontology;
KEGG, Kyoto Encyclopedia of Genes and Genomes; CDF, Cumulative
Distribution Function.
Table 2.
Results of GO and KEGG enrichment analysis.
ONTOLOGY ID Description p.adjust Count
BP GO:0007160 cell-matrix adhesion 6.21E-05 22
BP GO:0001933 negative regulation of protein phosphorylation 6.21E-05
31
BP GO:0042326 negative regulation of phosphorylation 0.000292 31
BP GO:0031589 cell-substrate adhesion 0.000304 26
BP GO:0007229 integrin-mediated signaling pathway 0.000693 13
BP GO:0070373 negative regulation of ERK1 and ERK2 cascade 0.000781 11
BP GO:0007369 gastrulation 0.001024 17
BP GO:0001503 ossification 0.00138 26
BP GO:0050808 synapse organization 0.00181 26
BP GO:1990778 protein localization to cell periphery 0.00181 22
CC GO:0031252 cell leading edge 8.26E-08 33
CC GO:0005911 cell-cell junction 8.26E-08 35
CC GO:0005925 focal adhesion 1.21E-07 32
CC GO:0005924 cell-substrate adherens junction 1.21E-07 32
CC GO:0030055 cell-substrate junction 1.23E-07 32
CC GO:0008305 integrin complex 1.41E-06 9
CC GO:0098636 protein complex involved in cell adhesion 2.95E-06 9
CC GO:0005913 cell-cell adherens junction 3.52E-06 15
CC GO:0030027 lamellipodium 4.17E-06 19
CC GO:0030175 filopodium 4.17E-06 14
MF GO:0050839 cell adhesion molecule binding 7.05E-11 44
MF GO:0045296 cadherin binding 3.17E-06 28
MF GO:0005178 integrin binding 0.000157 15
MF GO:0005543 phospholipid binding 0.000297 28
MF GO:0035091 phosphatidylinositol binding 0.000297 20
MF GO:0050840 extracellular matrix binding 0.000917 9
MF GO:0098631 cell adhesion mediator activity 0.001052 9
MF GO:0001968 fibronectin binding 0.003071 6
MF GO:0043236 laminin binding 0.004205 6
MF GO:0001618 virus receptor activity 0.004282 9
MF GO:0104005 hijacked molecular function 0.004282 9
KEGG hsa04512 ECM-receptor interaction 9.8E-06 14
KEGG hsa05412 Arrhythmogenic right ventricular cardiomyopathy 9.8E-06
13
KEGG hsa04510 Focal adhesion 9.62E-05 19
KEGG hsa05205 Proteoglycans in cancer 0.000399 18
KEGG hsa04810 Regulation of actin cytoskeleton 0.000705 18
KEGG hsa05165 Human papillomavirus infection 0.000705 23
KEGG hsa05410 Hypertrophic cardiomyopathy 0.000731 11
KEGG hsa05414 Dilated cardiomyopathy 0.001182 11
KEGG hsa04550 Signaling pathways regulating pluripotency of stem cells
0.002465 13
KEGG hsa04151 PI3K-Akt signaling pathway 0.003451 22
KEGG hsa04919 Thyroid hormone signaling pathway 0.007015 11
KEGG hsa05222 Small cell lung cancer 0.013198 9
KEGG hsa04929 GnRH secretion 0.02517 7
KEGG hsa05167 Kaposi sarcoma-associated herpesvirus infection 0.028658
13
KEGG hsa04390 Hippo signaling pathway 0.044293 11
KEGG hsa04310 Wnt signaling pathway 0.048091 11
KEGG hsa04935 Growth hormone synthesis, secretion and action 0.054994 9
KEGG hsa04925 Aldosterone synthesis and secretion 0.054994 8
KEGG hsa04750 Inflammatory mediator regulation of TRP channels 0.058929
8
KEGG hsa04611 Platelet activation 0.062936 9
[122]Open in a new tab
We then performed univariate Cox regression analysis to determine the
prognostic value of the 457 circadian rhythm subtype-related DEGs and
identified 127 genes associated with overall survival (OS). We
classified data into two gene clusters, A and B, using a consensus
clustering technique based on the expression patterns of the 127
OS-related DEGs in order to further confirm this regulation mechanism.
[123]Figure 6C displays the consensus matrix calculated through
consensus clustering, where k=2 successfully divides all samples into
two distinct groups. The color intensity represents the similarity
between samples, with darker colors indicating high similarity and
lighter colors indicating greater differences. This result lays a
foundation for subsequent analyses of potential biological functions.
([124] Figure 6D ) In the consensus matrix for k=3, the clustering of
samples appears more complex, suggesting the potential presence of
additional subgroups. Compared to k=2, the clustering quality shown
here is lower, indicating that k=3 may not be suitable for effective
sample classification. [125]Figure 6E shows the consensus CDF curves
for different k values. Observations reveal that the curve for k=2 has
the highest proportion in the region above 0.8, indicating better
clustering performance at this value. This further validates the
effectiveness of k=2 as the optimal choice. [126]Figure 6F indicates
that as the k value increases, the largest drop in the Delta area
occurs at k=2, supporting the use of k=2 for consensus clustering
analysis. This clearly reflects k=2 as the optimal number of clusters.
[127]Figure 6G shows the relationship between OS-related DEG expression
and clinicopathological features, with these genes being significantly
more expressed in gene cluster B than in gene cluster A. Kaplan-Meier
curves indicated that patients in gene cluster A had poorer OS, whereas
gene cluster B exhibited longer OS (P<0.001) ([128] Figure 6H ).
3.6. Construction of a circadian rhythm-related prognostic model
In this study, the training and validation sets used to construct the
circadian rhythm-related prognostic model were randomly allocated in a
1:1 ratio after merging the TCGA and GEO datasets. This process ensured
the representativeness of each dataset while providing a sufficient
sample size for model construction and validation. LASSO and
multivariate Cox regression analyses were conducted on the 127
OS-related DEGs to construct a prognostic model. Based on the smallest
partial likelihood deviation in LASSO regression analysis, we retained
13 genes: CMTM6, TNPO1, SLC5A3, CTBS, UTRN, HK2, SYNM, ZNF697,
CSGALNACT1, GBP3, LIF, TPD52L1, and BCAN ([129] Figures 7A, B ).
Multivariate Cox regression analysis ultimately produced a prognostic
model containing six genes, with the risk score calculated as follows:
[MATH:
Riskscore=CMT<
mi>M6*−0.286+TNPO1*0.523+
CTBS*−0.
361+UTRN*−0.209+HK2*
0.308+LIF*−
mo>0.131 :MATH]
. We found a significant difference in risk scores between gene
clusters A and B (P<2.2e-16), while no significant difference in risk
scores was observed between CRG clusters A and B ([130] Figures 7C, D
). [131]Figure 7E shows the relationship between molecular subtypes,
gene clusters, risk scores, and survival status. Kaplan-Meier survival
curves demonstrated that patients with low-risk scores had
significantly better OS than those with high-risk scores in the
training set, validation set, and overall dataset ([132] Figures 7F–H
). The ROC curves for the training set indicated that the model’s AUC
for 1, 3, and 5 years was 0.679, 0.664, and 0.661, respectively.
Additionally, ROC curves for the validation set and overall dataset
also demonstrated good predictive performance of the model ([133]
Figures 7I–K ).
Figure 7.
[134]Figure 7
[135]Open in a new tab
Construction of circadian rhythm prognostic model. (A, B) LASSO Cox
regression analysis of OS-DEGs. (C) Differences in risk scores between
gene cluster A and B. (D) Differences in risk scores between CRG
cluster A and B. (E) Sankey diagram showing the relationships between
molecular subtypes, gene clusters, risk scores, and survival status.
(F-H) Kaplan-Meier curves showing prognostic differences between high
and low-risk groups in the training set, validation set, and overall
dataset. (I-K) ROC curves of the training set, validation set, and
overall dataset. OS, Overall Survival; DEG, Differentially Expressed
Gene; LASSO, Least Absolute Shrinkage and Selection Operator; ROC,
Receiver Operating Characteristic Curve.
3.7. Gene Set Enrichment Analysis
To further explore the functions of the risk score, we performed GSEA
on the high-risk and low-risk groups. The results showed that the
high-risk group was primarily enriched in pathways such as cristae
formation, activation of the pre replicative complex, aminoacyl trna
biosynthesis, gluconeogenesis, and DNA replication, as well as
MYC_TARGETS_V2, hallmark myc targets v1, E2F targets, oxidative
phosphorylation, and G2M checkpoint.
In contrast, the low-risk group was mainly enriched in pathways such as
focal adhesion, GPCR ligand binding, G alpha I signaling events, toll
like receptor cascades, and apoptosis, as well as IL2 STAT5 signaling,
epithelial mesenchymal transition, KRAS signaling up, inflammatory
response, and interferon gamma response ([136] Figures 8A–D ,
[137]Table 3 ).
Figure 8.
[138]Figure 8
[139]Open in a new tab
GSEA enrichment analysis and immune infiltration analysis between
high-risk and low-risk groups. (A-D) GSEA enrichment analysis results
for high-risk and low-risk groups. (E) Correlation between risk scores
and immune cell infiltration levels. (F) Correlation between model
genes and immune cell infiltration levels. (G) Correlation of TME
scores between high-risk and low-risk groups. • indicates P<0.05, **
indicates P<0.01, *** indicates P<0.001. GSEA, Gene Set Enrichment
Analysis; TME, Tumor Microenvironment.
Table 3.
Results of GSEA enrichment analysis.
Description enrichmentScore p.adjust
HALLMARK_MYC_TARGETS_V2 0.766095 0.026483
HALLMARK_MYC_TARGETS_V1 0.56333 0.080515
HALLMARK_E2F_TARGETS 0.561358 0.080515
HALLMARK_OXIDATIVE_PHOSPHORYLATION 0.499392 0.080515
HALLMARK_G2M_CHECKPOINT 0.497002 0.080515
HALLMARK_IL2_STAT5_SIGNALING -0.64374 0.004145
HALLMARK_EPITHELIAL_MESENCHYMAL_TRANSITION -0.65837 0.004145
HALLMARK_KRAS_SIGNALING_UP -0.69475 0.004145
HALLMARK_INFLAMMATORY_RESPONSE -0.72604 0.004145
HALLMARK_INTERFERON_GAMMA_RESPONSE -0.77103 0.004145
KEGG_FOCAL_ADHESION -0.48112 0.019213
REACTOME_GPCR_LIGAND_BINDING -0.50436 0.019213
REACTOME_G_ALPHA_I_SIGNALLING_EVENTS -0.53569 0.019213
REACTOME_TOLL_LIKE_RECEPTOR_CASCADES -0.54188 0.019213
KEGG_APOPTOSIS -0.56615 0.019213
REACTOME_CRISTAE_FORMATION 0.847113 0.037178
REACTOME_ACTIVATION_OF_THE_PRE_REPLICATIVE_COMPLEX 0.742989 0.046003
KEGG_AMINOACYL_TRNA_BIOSYNTHESIS 0.692399 0.04434
REACTOME_GLUCONEOGENESIS 0.677523 0.049903
KEGG_DNA_REPLICATION 0.589663 0.049903
[140]Open in a new tab
3.8. Immune infiltration analysis between high-risk and low-risk groups
We assessed the relationship between risk scores and immune cell
abundance using the CIBERSORTx algorithm. As shown in [141]Figure 8E ,
risk scores were positively correlated with M0 macrophages and
negatively correlated with memory B cells, gamma delta T cells, and
activated CD4+ memory T cells. We also evaluated the relationship
between immune cell abundance and the six genes in the prognostic
model, finding that most immune cells, particularly HK2, were
significantly correlated with these genes (P<0.05) ([142] Figure 8F ).
We evaluated TME scores (stromal score, immune score, and ESTIMATE
score) between high-risk and low-risk groups using the ESTIMATE
package. We discovered that the high-risk group had significantly lower
stromal score (P<0.001), immune score (P<0.001), and ESTIMATE score
(P<0.001) than the low-risk group. ([143] Figure 8G ).
3.9. Prediction of immunotherapy response and drug sensitivity
We used TIDE scores to evaluate the effectiveness of immunotherapy. We
found that, compared to the low-risk group, patients in the high-risk
group had lower TIDE scores (P<0.001), lower Dysfunction scores
(P<0.001), and higher Exclusion scores (P=0.002) ([144] Figures 9A–C ).
According to the TIDE algorithm, immunotherapy responders had higher
risk scores (P<0.001) ([145] Figure 9D ). Kaplan-Meier curves indicated
that patients with both high-risk scores and high TIDE scores had the
worst prognosis, whereas those with low-risk scores and low TIDE scores
had the best prognosis (P<0.001) ([146] Figure 9E ). [147]Figure 9F
shows the relationships between TIDE scores, immunotherapy response,
risk scores, and survival status.
Figure 9.
[148]Figure 9
[149]Open in a new tab
Prediction of immunotherapy response and drug sensitivity. (A)
Differences in TIDE scores between high-risk and low-risk groups. (B)
Differences in Dysfunction scores between high-risk and low-risk
groups. (C) Differences in Exclusion scores between high-risk and
low-risk groups. (D) Differences in risk scores between immunotherapy
responders and non-responders as predicted by TIDE. (E) Kaplan-Meier
curve combining risk scores and TIDE scores. (F) Sankey diagram showing
the relationships between TIDE scores, immunotherapy response, risk
scores, and survival status. (G) Prediction of drug sensitivity between
high-risk and low-risk groups. TIDE, Tumor Immune Dysfunction and
Exclusion.
Next, we analyzed differences in chemotherapy drug sensitivity between
the high-risk and low-risk groups. The high-risk group exhibited
decreased sensitivity (increased IC50) to AKT inhibitor VIII
(P=0.00051), Camptothecin (P=3.1e-08), Gefitinib (P<0.001), Lapatinib
(P=3.6e-05), Nilotinib (P=1.2e-13), Rapamycin (P=0.00054), and
Temsirolimus (P=1.5e-11). In contrast, the sensitivity to Imatinib
(P=0.00097) was increased (decreased IC50) in the high-risk group
([150] Figure 9G ).
3.10. Construction of a nomogram scoring system
A univariate Cox regression analysis revealed a substantial correlation
between risk scores and a patient’s bad prognosis. (HR [95% CI] = 2.143
[1.589-2.892], P<0.001). Risk ratings for SKCM patients were found to
be an independent unfavorable prognostic factor by multivariate Cox
regression analysis. (HR [95% CI] = 2.165 [1.606-2.919], P<0.001)
([151] Figures 10A, B , [152]Table 4 ). Considering that the circadian
rhythm risk score is not convenient for clinical application in
predicting OS in SKCM patients, we developed a nomogram incorporating
risk scores and clinicopathological parameters to predict 1-year,
3-year, 5-year, and 10-year OS. The predictive factors included risk
score, patient age, and stage ([153] Figure 10C ). Calibration curves
demonstrated good predictive accuracy of the model ([154] Figure 10D ).
Time-dependent ROC results showed that the AUC values of the nomogram
in the training group for 1, 3, 5, and 10 years were 0.688, 0.737, and
0.690, respectively ([155] Figure 10E ).
Figure 10.
[156]Figure 10
[157]Open in a new tab
Construction of nomogram scoring system. (A) Results of univariate Cox
analysis of risk scores and clinicopathological factors. (B) Results of
multivariate Cox analysis of risk scores and clinicopathological
factors. (C) Construction of a Nomogram scoring system to predict
1-year, 3-year, and 5-year survival. (D) Calibration curve of the
Nomogram. (E) ROC curve of the Nomogram. ROC, Receiver Operating
Characteristic Curve.
Table 4.
Results of univariate and multivariate Cox analysis.
Characteristics Total(N) Univariate analysis Multivariate analysis
Hazard ratio (95% CI) P value Hazard ratio (95% CI) P value
Gender 413
Female 157 Reference
Male 256 1.036 (0.767-1.399) 0.819
Age 413
<=60 218 Reference
>60 195 1.622 (1.208-2.179) 0.001 1.664 (1.237-2.239) <0.001
Stage 413
I-II 224 Reference
III-IV 189 1.631 (1.218-2.185) 0.001 1.776 (1.324-2.380) <0.001
Riskscore 413
low 182 Reference
high 231 2.143 (1.589-2.892) <0.001 2.165 (1.606-2.919) <0.001
[158]Open in a new tab
4. Discussion
SKCM is recognized for its aggressive nature, high metastatic
potential, and considerable therapeutic challenges. According to the
latest cancer statistics, an estimated 97,610 new cases and 7,990
deaths are expected in the United States in 2023, reflecting a
continued upward trend in incidence ([159]1). This type of melanoma has
become one of the fastest-growing malignancies in Western countries,
particularly due to increased UV exposure and genetic predispositions.
Advancements in early detection and novel therapeutic strategies have
significantly improved survival rates for patients diagnosed at an
early stage of melanoma. However, the prognosis for patients with
late-stage or metastatic melanoma remains dismal. While non-metastatic
melanoma patients could expect a five-year survival rate of up to 99%,
this rate plummets to approximately 25% once the disease has
metastasized ([160]38). The molecular pathogenesis of SKCM is
intricate, involving multiple genetic alterations and signaling
pathways. Mutations in key oncogenes such as BRAF, NRAS, and c-KIT
activate pathways including MAPK/ERK and PI3K/AKT, which drive cellular
proliferation, survival, and migration. Approximately 48% of melanoma
cases harbor a BRAF mutation, with the BRAF V600E mutation being
particularly prevalent ([161]39). These mutations not only fuel tumor
growth but also contribute to the immune evasion mechanisms by
up-regulating PD-L1 expression and modulating the tumor
microenvironment to suppress immune surveillance ([162]40). In terms of
treatment, the advent of immune checkpoint inhibitors, such as
anti-PD-1 (nivolumab and pembrolizumab) and anti-CTLA-4 (ipilimumab)
antibodies, has revolutionized the management of advanced melanoma.
These therapies have significantly improved outcomes, with about 45% of
patients experiencing durable responses. However, the treatment success
is not uniform, with resistance developing in a substantial subset of
patients ([163]41, [164]42). Moreover, the management of immune-related
adverse events, which could significantly impact the quality of life,
remains a critical aspect of treatment strategies.
Circadian Rhythm Genes (CRGs) orchestrate a vast array of biological
processes vital for organismal homeostasis, including the cell cycle,
DNA repair mechanisms, and metabolic pathways, all critical in
tumorigenesis and cancer progression. The core clock mechanism involves
a transcription-translation feedback loop wherein CLOCK and BMAL1
heterodimers promote the expression of period (Per1/2/3) and
cryptochrome (Cry1/2) genes. These proteins, in turn, regulate their
expression by interacting with CLOCK and BMAL1, establishing a
circadian rhythm that impacts various cellular functions ([165]43,
[166]44). In the context of oncology, disruptions in CRG expression
notably influence tumor behavior, affecting proliferation, apoptosis,
and metastasis across several cancer types. For instance, in breast
cancer, altered expression of Per genes has been linked to increased
cell proliferation and reduced apoptosis, thereby exacerbating cancer
progression. Similarly, studies in colorectal cancer have identified
mutations in various CRGs that correlate with tumor growth and poor
patient prognosis due to disrupted circadian control over cell cycle
and apoptosis ([167]45, [168]46).
In SKCM, recent studies have elucidated that CRGs such as CLOCK and
BMAL1 significantly influence various cellular functions including the
tumor microenvironment, cell cycle, apoptosis regulation, DNA damage
response, metabolic reprogramming, and immune evasion, thus affecting
the response to therapy. These genes modulate the functionality of
immune cells within the tumor microenvironment, potentially altering
immunotherapy outcomes. For example, the circadian clock regulates the
timing of UV exposure, which impacts the efficacy of DNA repair
mechanisms and consequently influences melanoma risk and progression
([169]47, [170]48). Furthermore, CRGs interact directly with key cell
cycle and apoptosis regulators. Studies have demonstrated that CLOCK
and BMAL1 affect cell proliferation by regulating the expression of
Wee1, a kinase that inhibits cell cycle progression from G2 to M phase,
suggesting a mechanism where disruptions in circadian rhythms could
promote uncontrolled cell growth typical of melanoma ([171]49).
Additionally, the timing of DNA repair processes is controlled by
circadian regulators, which modulate the expression and activity of
various nucleotide excision repair enzymes. In SKCM, where UV-induced
DNA damage is a critical risk factor, the efficiency of repair
mechanisms during peak UV exposure times could significantly influence
the risk of mutation accumulation and tumor initiation. Research has
shown that the expression of XPA, a crucial DNA repair protein, is
circadian-regulated, aligning DNA repair processes with periods of
likely UV exposure ([172]50). Moreover, melanoma cells exhibit unique
metabolic profiles influenced by the circadian clock. CRGs like BMAL1
are involved in the regulation of oxidative phosphorylation and
glycolysis pathways, which are often altered in cancer cells to meet
their increased energy demands. Disruption of these pathways could
create a metabolic environment that facilitates melanoma progression
and resistance to therapy ([173]51).
Our study rigorously explored the genetic and transcriptional
landscapes of CRGs in SKCM using advanced bioinformatics tools. By
analyzing comprehensive datasets from TCGA and GEO, we identified
significant somatic mutations and CNVs in CRGs such as RORB, PER3, and
CLOCK, which are associated with diverse clinical outcomes and immune
infiltration patterns in SKCM. Upon comparing mRNA expression levels of
CRGs in tumor and normal tissues, it was observed that the expression
of the TIMELESS gene was elevated and associated with poor prognosis,
confirming findings from previous studies. Zhao et al. have
specifically noted that TIMELESS may regulate DNA replication and cell
cycle-related genes, potentially influencing melanoma progression
([174]52). In contrast, RORA expression was notably reduced in tumor
tissues. Consistent with findings from the Benna study, this reduction
and the presence of rs339972 C and rs10519097 T alleles of RORA were
linked to a decreased risk of developing melanoma ([175]53). Our
prognostic model suggests that overexpression of PER predicts a poor
prognosis in melanoma. The implicated mechanism involves the
recognition of m6A-modified PER1 by the protein YTHDF2, which may
accelerate melanoma progression ([176]54). More importantly, using
LASSO regression analysis to identify the smallest partial likelihood
deviation, followed by multivariate Cox regression analysis, we
ultimately developed a prognostic model comprising six genes. These
genes are believed to play critical roles in the progression of CMM.
CMTM6 has been identified as a key regulator of immune evasion in tumor
cells. It promotes tumor cell escape from immune surveillance by
modulating the expression of immune checkpoint molecules, particularly
programmed death-ligand 1 (PD-L1). CMTM6 stabilizes PD-L1 on the
surface of tumor cells by preventing its lysosomal degradation, thereby
enhancing the tumor’s ability to suppress T cell-mediated immune
responses in the tumor microenvironment. This mechanism underscores the
pivotal role of CMTM6 in the regulation of immune checkpoint pathways,
which are critical in tumor immune evasion and resistance to
immunotherapy ([177]55, [178]56). CMTM6 also influences the interaction
between tumor cells and immune cells within the tumor microenvironment.
Studies suggest that CMTM6 may modulate the activity of
tumor-associated macrophages (TAMs), natural killer (NK) cells, and
dendritic cells, thereby shaping the immune landscape of the tumor
microenvironment. For instance, CMTM6-mediated stabilization of PD-L1
can inhibit the activation of cytotoxic NK cells and impair the
antigen-presenting function of dendritic cells, further promoting an
immunosuppressive microenvironment conducive to tumor growth and
progression ([179]57). The clinical significance of CMTM6 lies in its
potential as a prognostic biomarker and therapeutic target. Elevated
expression of CMTM6 has been associated with poor prognosis, including
reduced survival rates and increased risk of tumor recurrence in
several cancer types, such as lung cancer, melanoma, and colorectal
cancer ([180]58–[181]60). Targeting CMTM6 has been proposed as a novel
therapeutic strategy to enhance the efficacy of immune checkpoint
blockade therapies by destabilizing PD-L1 and restoring T cell-mediated
antitumor immunity. These findings highlight modulating CMTM6
expression may offer new avenues for cancer immunotherapy, particularly
in tumors where PD-L1 plays a critical role in immune evasion
([182]59). TNPO1 is a transport protein involved in nuclear-cytoplasmic
trafficking which plays a critical role in regulating the cell cycle
and is implicated in tumor cell proliferation and metastasis.
Overexpression of TNPO1 has been linked to increased tumor cell
proliferation and migration, contributing to aggressive cancer
phenotypes. For instance, elevated TNPO1 expression has been observed
in colorectal cancer, where it correlates with enhanced tumor growth
and poor prognosis ([183]61). TNPO1 influences tumor progression
through its involvement in critical oncogenic signaling pathways. It
has been shown to regulate components of the PI3K/Akt and MAPK
pathways, which are essential for tumor cell survival, proliferation,
and migration. By modulating the nuclear transport of signaling
proteins, TNPO1 indirectly affects downstream signaling cascades that
drive tumor growth and metastasis ([184]61). TNPO1 is a key mediator of
nucleocytoplasmic transport, specifically facilitating the nuclear
import of RNA and proteins critical for cellular homeostasis. In tumor
cells, TNPO1 plays a pivotal role in the transport of transcription
factors, splicing regulators, and other nuclear-localized proteins that
govern gene expression and cellular behavior. This function is
particularly relevant in cancer, as dysregulated nuclear transport can
lead to aberrant gene expression patterns that promote tumor growth,
invasion, and metastasis ([185]62). Additionally, TNPO1-mediated
transport may contribute to the adaptation of tumor cells to their
microenvironment, enhancing their ability to evade immune surveillance
and thrive in metastatic sites ([186]62, [187]63). Moreover, TNPO1 has
been linked to melanoma cell proliferation and metastasis through its
regulation of nuclear import/export processes. Its dysregulation is
associated with poor prognosis in cancers ([188]64). Cathepsin B serine
protease is involved in extracellular matrix remodeling and tumor
invasion. Elevated expression of CTBS facilitates melanoma metastasis
by promoting cell migration and invasiveness ([189]65). A key
glycolytic enzyme, hexokinase 2 (HK2) is associated with metabolic
reprogramming in melanoma cells. Its upregulation promotes the Warburg
effect, which supports tumor growth and survival ([190]66). SLC5A3
plays a central role in ion transport and energy metabolism. Its
expression levels may directly impact the metabolic state of cells,
thereby influencing tumor growth and metastatic potential ([191]67).
These investigations extend beyond previous studies by not only
identifying these variations but also linking them to functional
consequences within the tumor microenvironment. The methodological
approach of this study stands out due to its use of consensus
clustering to classify SKCM patients into distinct molecular subtypes,
CRG cluster A and B, based on CRG expression profiles.
The GSVA analysis results revealed that CRG cluster B was significantly
enriched with several pathways closely associated with cellular
biological functions. For example, in Protein Synthesis and Metabolism:
the
“GOBP_POSITIVE_REGULATION_OF_RESPONSE_TO_ENDOPLASMIC_RETICULUM_STRESS”
pathway is involved in regulating the cellular response to endoplasmic
reticulum (ER) stress. This pathway may be linked to the survival and
adaptability of melanoma cells under external stress conditions
([192]68). What’s more, Transport and Signal Transduction: The
“KEGG_MTOR_SIGNALING_PATHWAY” is closely associated with cell growth,
proliferation, and survival. This pathway is a critical regulator of
tumor cell metabolism and drug resistance, playing a pivotal role in
melanoma progression ([193]69). Correspondingly, recent research has
corroborated that selective inhibition of these pathways(mTOR and ERBB)
presents a promising therapeutic avenue for melanoma variants that
exhibit resistance to PD-1 monoclonal antibodies, which are critical in
the disease’s advancement ([194]70, [195]71). Biosynthesis and Repair
Pathways: Pathways such as “GOBP_GOLGI_ORGANIZATION” and
“GOBP_MAGNESIUM_ION_TRANSPORT” suggest that alterations in endoplasmic
reticulum and Golgi apparatus functions may support the biosynthetic
and repair processes required for tumor growth in melanoma cells
([196]72). Moreover, The GSEA analysis results revealed significant
differences in biological processes and signaling pathways between the
high-risk and low-risk groups. Pathways Enriched in the High-Risk
Group: The high-risk group was primarily enriched in pathways such as
“REACTOME_ACTIVATION_OF_THE_PRE_REPLICATIVE_COMPLEX” and
“KEGG_DNA_REPLICATION.” These findings suggest that enhanced cell cycle
regulation and DNA replication in melanoma cells may contribute to the
rapid proliferation of tumor cells. These pathways provide potential
therapeutic targets, particularly for the use of cell cycle inhibitors
in cancer treatment ([197]73–[198]75). Pathways Enriched in the
Low-Risk Group: The low-risk group exhibited enrichment in pathways
such as “KEGG_FOCAL_ADHESION” and “REACTOME_GPCR_LIGAND_BINDING.” This
indicates that patients in the low-risk group may have stronger
cell-cell interactions and adhesion abilities. These findings imply
that tumors in the low-risk group are more effectively integrated
within the tumor microenvironment, potentially linked to local immune
responses or tumor suppression ([199]76, [200]77). Targeting these
pathways could potentially disrupt the melanoma’s progression
mechanisms, offering new hope for treatment-resistant forms of this
malignancy. Furthermore, the immune infiltration analysis demonstrated
that CRG cluster A had higher infiltration of immune cells such as
activated CD8+ T cells and macrophages, which are known to be critical
for anti-tumor immunity, compared to CRG cluster B. The differential
immune infiltration between high-risk and low-risk groups provides
insights into the prognostic value of CRGs in SKCM. High-risk patients,
as identified by our prognostic model, exhibited lower levels of immune
cell infiltration, which might contribute to their poorer prognosis.
Conversely, low-risk patients showed higher immune scores, indicating a
more robust immune surveillance mechanism. In our study, we observed
significant differences in immune infiltration between the high-risk
and low-risk groups, which may have important implications for the
effectiveness of immunotherapy. In the high-risk group, lower levels of
immune cell infiltration may reflect a microenvironment characterized
by immune evasion mechanisms, such as the upregulation of immune
checkpoint molecules or the recruitment of immunosuppressive cells,
such as regulatory T cells (Tregs) and myeloid-derived suppressor cells
(MDSCs). These mechanisms are known to diminish immune system activity,
potentially leading to reduced responses to immunotherapy and weakened
immune surveillance, allowing tumor cells to evade recognition by the
immune system ([201]78, [202]79). The immune microenvironment in the
high-risk group suggests that these patients may respond poorly to
conventional immune checkpoint inhibitors (e.g., PD-1/PD-L1
inhibitors). This is consistent with studies that have demonstrated
that immunosuppressive microenvironments reduce the efficacy of
immunotherapy. Furthermore, the increased presence of MDSCs may
exacerbate immune suppression by secreting immunosuppressive cytokines,
thereby further impairing anti-tumor immune responses and enhancing
tumor immune evasion. In contrast, the low-risk group exhibits higher
immune cell infiltration, which may indicate a more active and
effective anti-tumor immune response and likely to support more robust
anti-tumor responses, making these patients more sensitive to
immunotherapy. Elevated levels of effector T cells and natural killer
(NK) cells suggest stronger immune surveillance, which has been
associated with improved responses to immune checkpoint blockade
therapies ([203]80). These findings underscore the potential for
stratifying patients based on immune infiltration patterns to predict
immunotherapy outcomes and optimize treatment strategies. Further
analysis of these differences could enhance our understanding of the
tumor immune microenvironment and guide the development of combination
therapies that target immune evasion mechanisms in high-risk patients,
improving their response to treatment. This finding aligns with the
immunosuppressive nature of the TME in high-risk groups and suggests
that CRG expression could be a key modulator of immune evasion in
melanoma ([204]81). CRGs such as CLOCK and BMAL1 play pivotal roles in
modulating the tumor immune microenvironment, influencing immune cell
function and potentially contributing to immune evasion mechanisms. The
CLOCK protein, a core component of the circadian clock, forms a
heterodimer with BMAL1 to regulate the expression of genes involved in
various physiological processes, including immune responses. Disruption
of CLOCK expression can lead to altered immune cell function, impacting
the body’s ability to mount effective anti-tumor responses. Studies
have shown that circadian disruptions can affect the recruitment and
function of myeloid-derived suppressor cells (MDSCs), which facilitate
immune evasion in tumors ([205]82). On the other hand, BMAL1 is
integral to maintaining the circadian rhythm and has been implicated in
the regulation of immune cell metabolism and function. In macrophages,
BMAL1 acts as a metabolic sensor, influencing inflammatory responses
and phagocytic activity. Deficiency in BMAL1 has been associated with
reduced inflammatory responses and altered metabolic processes in
microglial cells, suggesting its role in modulating immune cell
activity within the tumor microenvironment ([206]83). Disruptions in
circadian clock genes like CLOCK and BMAL1 can lead to immune
dysregulation, creating an environment conducive to tumor progression.
Abnormal circadian rhythms have been linked to upregulation of immune
inhibitory molecules such as PD-L1 and CTLA-4, contributing to T cell
exhaustion and immune evasion in cancer ([207]84). According to Adegoke
et al., patients whose tumors had a higher infiltration of immune
cells, such as macrophages, experienced better response rates and
longer progression-free survival compared to those with tumors
characterized by moderate or scarce immune cell presence ([208]81).
Future research should focus on elucidating the precise mechanisms by
which these CRGs influence immune cell infiltration and function in
SKCM, potentially uncovering novel therapeutic targets to enhance
anti-tumor immunity. The predictive accuracy of our risk model was
further validated using ROC curve analysis, with the nomogram
integrating clinical parameters and risk scores demonstrating
substantial predictive power for 1-, 3-, and 5-year overall survival.
Compared to other studies, such as those by Cabrita and Cirenajwis
([209]14, [210]15), which predominantly focused on broader genomic
profiling, our research specifically targets the circadian regulatory
network. This unique focus allows us to provide novel insights into the
temporal dynamics of gene expression in SKCM. Unlike general genomic
studies that provide a static snapshot of genetic alterations, our
approach delves into how these alterations are influenced by the body’s
circadian rhythms, offering a dynamic perspective that is critical for
understanding tumor behavior over time. In our drug sensitivity
analysis, the sensitivity differences to AKT inhibitor VIII were
particularly notable. Although AKT inhibitors have not yet become a
standard part of melanoma treatment, their clinical potential should
not be overlooked. The PI3K/AKT signaling pathway plays a crucial role
in melanoma progression, making AKT a promising therapeutic target.
Several AKT inhibitors have been developed and are undergoing clinical
evaluation. For instance, ipatasertib, an oral AKT inhibitor, has shown
potential in clinical trials targeting tumors with specific genetic
alterations. Additionally, the combination of AKT inhibitors with other
therapeutic agents is being explored to overcome resistance mechanisms
and enhance treatment efficacy. For example, combining AKT inhibitors
with BRAF inhibitors has demonstrated synergistic effects in
preclinical melanoma models, suggesting a potential strategy for
patients with BRAF-mutant melanoma ([211]85). As a result, AKT
inhibitors have shown potential therapeutic value in inhibiting tumor
cell proliferation and promoting apoptosis ([212]86). Currently,
multiple AKT inhibitors are undergoing clinical trials to explore their
specific applications in melanoma treatment. Among them, AKT inhibitor
VIII has demonstrated efficacy in selectively targeting the AKT pathway
and effectively suppressing melanoma cell proliferation, providing a
promising avenue for personalized treatment strategies. Camptothecin, a
topoisomerase I inhibitor, has shown anti-tumor activity in various
cancers, including melanoma ([213]87). However, due to its toxicity and
solubility issues, derivatives such as irinotecan and topotecan have
been developed and are currently used in clinical settings. These
agents interfere with DNA replication in rapidly dividing cells,
leading to cell death ([214]88). While not standard treatments for
melanoma, camptothecin derivatives are being investigated in clinical
trials, either alone or in combination with other therapies, to assess
their efficacy and safety profiles in melanoma patients ([215]89). Our
findings suggest that patients in the low-risk group may exhibit
increased sensitivity to AKT inhibitors and camptothecin derivatives,
potentially leading to better therapeutic outcomes. Conversely,
high-risk patients may require alternative strategies or combination
therapies to achieve optimal results. Incorporating these insights into
clinical decision-making could enhance personalized treatment
approaches for melanoma patients.
In this study, we performed differential expression analysis between
tumor and normal groups using the limma package, a widely accepted and
well-performing method that has been successfully applied in many
high-throughput data analyses. Furthermore, our analysis was primarily
focused on key biological hypotheses, and the differentially expressed
genes identified have been partially validated in the literature
([216]49, [217]53). Therefore, the primary objective of this study is
to investigate the impact of circadian rhythms on tumorigenesis and
progression, identify key genes associated with SKCM and their
potential mechanisms, and develop a prognostic model. SNPs and CNVs, as
genetic markers, can influence patients’ clinical manifestations
through various mechanisms ([218]90). These variations can lead to
changes in gene function, thereby affecting disease progression and
treatment responses at multiple levels. In SKCM patients, CNVs may
result in the loss of key tumor suppressor genes or the activation of
oncogenes, thus promoting tumor development. For example, the loss of
tumor suppressor genes can disrupt cell cycle checkpoints, leading to
faster cell proliferation ([219]91, [220]92). SNPs may affect the
function of key genes in signaling pathways, altering cell growth
regulation and impacting tumor proliferation and progression ([221]93).
Certain SNPs and CNVs can modify the structure and function of drug
targets or drug-metabolizing enzymes, directly influencing patients’
responses to specific drugs ([222]94). Genetic variations can also
affect immune cell infiltration, angiogenesis, and other factors,
thereby altering the tumor microenvironment and influencing patient
survival rates ([223]95). Furthermore, our study lays the groundwork
for more personalized approaches in cancer treatment, where circadian
biomarkers could predict patient-specific optimal treatment times,
enhancing survival rates and quality of life for melanoma patients.
Although our study provides robust insights, it is not without
limitations. The reliance on retrospective data and potential biases in
the TCGA and GEO datasets could affect the generalizability of the
findings. Moreover, functional validation of CRG-related mechanisms in
experimental models is necessary to confirm the causative roles
suggested by our analyses. Although the prognostic model we constructed
demonstrates good predictive performance, the lack of external dataset
validation may limit the generalizability and applicability of the
results. Future research will consider incorporating independent
datasets to enhance the comprehensiveness of the findings and we plan
to collect clinical samples and employ experimental methods such as
qRT-PCR or IHC to further validate the conclusions of this study. Last
but not least, immune infiltration analysis was included as a
complementary aspect to broaden the research perspective and provide
insights into the role of circadian rhythm-related genes in skin
cutaneous melanoma. It is critical to incorporate additional
experimental validations in future studies to enhance the credibility
and comprehensiveness of our research.
Our study’s integration of circadian genomics with immune profiling
offers a groundbreaking perspective, suggesting that the timing of
therapeutic interventions could be crucially optimized based on
circadian influences. This innovative approach not only proposes the
possibility of enhancing treatment efficacy but also suggests reducing
side effects by aligning treatment schedules with the body’s biological
clock. Such synchronization could exploit the natural peak activity
phases of cancer cell vulnerability and immune system responsiveness,
potentially transforming the strategic planning of melanoma therapy.
The implications of our findings extend beyond the immediate benefits
of therapy timing. By understanding the circadian modulation of gene
expression, our research significantly contributes to the broader
understanding of melanoma pathogenesis. This knowledge is pivotal for
developing future interventions that are finely tuned to the biological
rhythms of the body, thereby improving prognostic assessments and
therapeutic outcomes.
5. Conclusion
This study systematically explored the role of CRGs in SKCM. Our
analysis identified significant gene alterations and differential
expression patterns of CRGs between tumor and normal tissues, revealing
their association with poor prognosis and immune modulation. Two
molecular subtypes were established based on CRG expression, showing
distinct clinical features and immune infiltration patterns. A
prognostic model comprising six key CRGs (e.g., CMTM6, TNPO1, and HK2)
demonstrated reliable accuracy in predicting overall survival.
High-risk scores were linked to elevated immune checkpoint expression,
decreased immune cell infiltration, and lower sensitivity to
chemotherapies, underscoring CRGs’ impact on immune evasion and therapy
resistance.
Acknowledgments