Abstract
Among women, breast carcinoma is one of the most complex cancers, with
one of the highest death rates worldwide. There have been significant
improvements in treatment methods, but its early detection still
remains an issue to be resolved. This study explores the multifaceted
function of hyaluronan-mediated motility receptor (HMMR) in breast
cancer progression. HMMR’s association with key cell cycle regulators
(AURKA, TPX2, and CDK1) underscores its pivotal role in cancer
initiation and advancement. HMMR’s involvement in microtubule assembly
and cellular interactions, both extracellularly and intracellularly,
provides critical insights into its contribution to cancer cell
processes. Elevated HMMR expression triggered by inflammatory signals
correlates with unfavorable prognosis in breast cancer and various
other malignancies. Therefore, recognizing HMMR as a promising
therapeutic target, the study validates the overexpression of HMMR in
breast cancer and various pan cancers and its correlation with certain
proteins such as AURKA, TPX2, and CDK1 through online databases.
Furthermore, the pathways associated with HMMR were explored using
pathway enrichment analysis, such as Gene Ontology, offering a
foundation for the development of effective strategies in breast cancer
treatment. The study further highlights compounds capable of inhibiting
certain pathways, which, in turn, would inhibit the upregulation of
HMMR in breast cancer. The results were further validated via MD
simulations in addition to molecular docking to explore
protein–protein/ligand interaction. Consequently, these findings imply
that HMMR could play a pivotal role as a crucial oncogenic regulator,
highlighting its potential as a promising target for the therapeutic
intervention of breast carcinoma.
Keywords: HMMR, breast cancer, biomarkers, Cdk1, AURKA, Tpx2, mTOR, MD
simulation
Graphical Abstract
[44]graphic file with name FPHAR_fphar-2024-1361424_wc_abs.jpg
Highlights
* • The glycoprotein hyaluronan-mediated motility receptor (HMMR),
encoded by the HMMR gene on chromosome 5, exhibits a spiral
structure and serves diverse functions in cell growth, cancer
metastasis, cellular pluripotency, and resistance to cancer
treatments.
* • Although HMMR expression is carefully regulated in healthy
tissues, it undergoes upregulation in proliferative tissues,
contributing to the invasive nature and metastasis observed in
various carcinomas, including breast, colorectal, and prostate
carcinomas.
* • HMMR is associated with the cell cycle, particularly in the S/G2
and G2/M phases, by potentially impacting CDK1 and AURKA levels and
also the pivotal PI3K/AKT/mTOR pathway that is essential for cell
division. It plays a significant role in the G2/M phase by directly
phosphorylating CDK1 activators and inhibitors. Due to persistent
HMMR overexpression in cancer, it emerges as a promising
therapeutic target. Inhibiting HMMR expression through compounds
targeting key signaling pathways, such as mTOR, presents innovative
therapeutic avenues, particularly in breast cancer.
* • Our study underscores the multifaceted involvement of HMMR in
carcinoma initiation and advancement, emphasizing its importance as
a therapeutic target in breast cancer. Our study proposes specific
inhibitors for further investigation in breast cancer treatment.
1 Introduction
Despite extensive exploration in both basic and clinical research,
along with trials of potential innovative treatments, cancer persists
as a significant global health challenge ([45]Siegel et al., 2023). As
per GLOBOCAN (2020) data, breast cancer continues to be one of the most
frequently occurring cancers. Manifesting as a diverse malignancy, it
constitutes as the foremost factor in cancer-related fatalities among
women ([46]Stickeler, 2011). The detection of molecular biomarkers that
can function as prognostic and predictive markers has aided healthcare
professionals in therapeutic choices. This allows for the application
of a more individualized approach to treatment, optimizing therapy, and
averting the scenarios of excessive treatment, insufficient treatment,
and inaccurate treatment ([47]Stickeler, 2011; [48]Mir et al., 2023).
Hyaluronan-mediated motility receptor (HMMR), also known as CD168 or
RHAMM (receptor for hyaluronic acid-mediated motility), is a
glycoprotein with spiral configuration encoded by the HMMR gene
situated on the human chromosome 5 (5q33.2-qter) ([49]Tolg et al.,
2010). As per the available information, HMMR displays a significant
coiled-coil (CC) structure containing numerous binding sites for its
associates ([50]Hardwick et al., 1992; B; [51]Yang et al., 1994). In
the beginning, HMMR was identified as an innovative hyaluronan-mediated
mobility receptor and a microtubule-associated spindle assembly factor
([52]Hofmann et al., 1998). However, extensive research now suggests
that HMMR performs multiple functional roles in overseeing cell growth,
the spread of cancer to other parts of the body (D. [53]Yang et al.,
2021), the preservation of cellular pluripotency ([54]Tilghman et al.,
2014), and resistance to cancer treatment ([55]Sofi et al., 2023; H;
[56]Zhang et al., 2019) in various malignancies, such as lung carcinoma
([57]Li W. et al., 2020), hepatic neoplasia (D. [58]Zhang et al.,
2020), urothelial carcinoma (D. [59]Yang et al., 2021), and stomach
neoplasia ([60]Kang et al., 2020). Studies have revealed that the
regulation of HMMR expression is closely managed in healthy tissues,
yet it undergoes upregulation in proliferative tissues. This
contributes to invasiveness and metastasis, leading to an unfavorable
prognosis in various human carcinomas, including colorectal ([61]Zlobec
et al., 2008), breast ([62]Assmann et al., 2001), and prostate
carcinomas ([63]Gust et al., 2009).
HMMR exhibits diverse functions, which vary based on whether it is
located extracellularly or intracellularly. In the extracellular
capacity, HMMR functions as extensively characterized receptors for
hyaluronic acid (HA), regulating cell migration induced by HA, a
pivotal aspect of both inflammation and the recuperative process during
wound healing associates ([64]Hardwick et al., 1992; B; [65]Yang et
al., 1994). In the intracellular capacity, it serves as a protein
associated with the cell cycle, overseeing the formation of the mitotic
spindle and microtubules ([66]Assmann et al., 2001). Significantly,
these inherent roles of HMMR are frequently disrupted in cancer,
leading to growth benefits and advancements in the disease (Y.-T.
[67]Chen et al., 2018). HMMR expression in most healthy tissues at a
balanced state is generally minimal, and its expression is triggered in
response to inflammatory signals ([68]Tolg et al., 2021). On the other
hand, human neoplasms have been reported to display increased
concentrations of HMMR, and this upregulation is often associated with
metastatic tendencies, formidable traits, and an adverse prognosis in
instances of prostate and hematological carcinomas (Y.-T. [69]Chen et
al., 2018). The excessive expression of HMMR also operates as a
standalone predictor of prognosis in many cancer types ([70]Mele et
al., 2017; [71]Song et al., 2018).
HMMR shows potential to indirectly aid in the microtubule assembly by
facilitating the precise positioning of TPX2, thereby promoting the
efficient activation of aurora kinase at specific centers crucial for
microtubule formation (H. [72]Chen et al., 2014). HMMR serves as a
collaborative protein with TPX2. The trilateral association of these
two proteins, along with dynein, upholds the structural integrity of
the spindle and enhances the concentration of spindle poles
([73]Maxwell et al., 2003). TPX2 acts as a co-activator of AURKA and
plays a dual role: it alone can boost kinase activity beyond its
baseline levels and is indispensable for achieving optimal kinase
functionality. Given that approximately 40%–60% of TPX2 forms a complex
with HMMR during mitosis in human cells, HMMR emerges as a significant
contributor of the cellular processes. HMMR does not possess a
transmembrane domain, allowing it to associate with various receptors
embedded in the membrane, such as CD44, and also transforming growth
factor β1 (TGF-β1) ([74]Bayliss et al., 2003). This interaction
triggers the intracellular signaling pathway related to cellular
migration. A mechanism by which HMMR might enhance cellular
multiplication is by amplifying the function of M-phase-encouraging
factor CDK1 activity through the fortification of Cdc2 mRNA levels in a
hyaluronic acid-dependent manner ([75]Mohapatra et al., 1996). The
CDK1/cyclin B complex governs the initiation of mitosis by
phosphorylating diverse spindle-related proteins, such as importin, to
regulate the efficient formation of microtubules and segregation of
chromosomes ([76]Guo et al., 2019). Consequently, the overexpression of
HMMR may expedite the premature initiation of mitosis, resulting in a
genetic modification that supports the growth of tumors through an
upsurge in CDK1 activity.
Cell cycle advancement via the S/G2 and G2/M phases in both meiosis and
mitosis relies on the vital activation of Cdk1, a catalytic unit of the
M phase-promoting factor (MPF). During the G2/M phase, the mTOR pathway
assumes a crucial role in facilitating cell division by directly
phosphorylating the activators and inhibitors of CDK1. This underscores
the importance of AKT in regulating the cell cycle and highlights the
importance of the PI3K/AKT pathway in promoting cellular division. Due
to the persistent overexpression of HMMR in cancer and its manifold
functions in cancer initiation and advancement, it holds substantial
promise as a target for cancer treatment. Consequently, the
identification of substances that obstruct HMMR expression in tumors
may present innovative therapeutic approaches to curb tumor growth.
Here, we demonstrate HMMR and its correlation with certain potential
genes such as AURKA, TPX2, and CDK1 that are involved in the cell cycle
and are useful for the activation of certain pathways, which may be
targeted to reduce HMMR overexpression. We have further tried to
demonstrate how HMMR can be inhibited through certain compounds that
are key regulators of important signaling pathways which aid in breast
cancer progression. It was seen through different databases that HMMR
was highly upregulated in various malignancies, including breast
cancer, and HMMR has a correlation with AURKA, TPX2, and CDK1, which
are also involved in the upregulation of mTOR signaling pathways.
Therefore, in this study, we propose that targeting mTOR through
certain inhibitors like rapamycin and Torin2 is of great significance
as targeting the mTOR pathway will inhibit HMMR overexpression due to a
strong correlation between HMMR and the genes involved in mTOR and the
cell cycle pathway (AURKA, TPX2, and CDK1). Hence, HMMR holds a great
therapeutic significance.
2 Methodology
2.1 Predicting a potential target gene associated with the respective disease
(breast cancer)
The Gene Cards Database ([77]https://www.genecards.org/) ([78]Safran et
al., 2010), OMIM ([79]https://www.omim.org/) ([80]Hamosh et al., 2000),
and STITCH ([81]https://www.omim.org/) ([82]Kuhn et al., 2010) were
utilized to retrieve extensive and user-friendly details regarding
genes that are either anticipated or established as potential
biomarkers linked with breast cancer. These amalgamated databases were
utilized to collect information on diverse genes associated with breast
cancer. The term “breast cancer” was applied as a primary keyword in
this investigation.
2.2 UALCAN
UALCAN serves as a comprehensive online data repository for the
exploration of cancer omics data ([83]Chandrashekar et al., 2017).
Developed using PERL-CGI, it offers high-quality graphs and box plots,
facilitating convenient access to publicly accessible cancer OMICS data
from sources such as TCGA, CPTAC, and CBTTC. The platform aids in the
identification of potential biomarkers for the in silico validation of
selected genes, and it provides comprehensive expression profiles for
these genes across various carcinomas. Moreover, UALCAN offers valuable
patient survival information for protein-coding genes, as well as
miRNAs and lncRNAs. To investigate the impact of HMMR expression on
various cancer types, data from the UALCAN dataset were consulted.
2.3 TIMER 2.0
A thorough tool for methodologically analyzing the expression and
estimate of immunological filtrates across various carcinomas is TIMER
2.0 ([84]http://timer.comp-genomics.org/) ([85]Li T. et al., 2020).
Methods like CIBERSORT, quanTIseq, MCPcounter, and EPIC have made it
possible to observe the correlation between immune cell types, gene
expression, and mutation status in the database. To gain insights into
the gene expression patterns of HMMR, a heat map of the protein was
created using TIMER 2.0 across all malignancies.
2.4 Constructing a shared network that connects the disease biomarkers
A search of the STRING database ([86]https://stringdb.org/)
([87]Szklarczyk et al., 2021) may be used to identify important
regulatory genes implicated in the illness and look for information on
protein–protein interactions (PPIs) ([88]Lehne and Schlitt, 2009). It
consists of a plethora of data regarding known and predicted
protein–protein interactions of several species ([89]Szklarczyk et al.,
2016). PPIs were identified using STRING, with a minimum confidence
level of >0.4 as the cutoff value. Only “Homo sapiens” was included in
the inquiry, and high-confidence ratings of more than 0.7 were not
shared with STRING prior to the verified targets being sent there. In
the end, PPI data were located again. The primary targets for treating
breast cancer were determined to be the top proteins with the greatest
levels of expression.
2.5 bc-GenExMiner
The annotated BC transcriptomic and RNA-Seq data have been made
accessible online via the Breast Cancer Gene-Expression Miner V4.5
([90]http://bcgenex.ico.unicancer.fr/) (!!! INVALID CITATION !!!
([91]Jézéquel et al., 2012; [92]Jezequel et al., 2013)). It was applied
to investigate the relationship between HMMR expression levels and
certain clinical characteristics of individuals with breast tumors. The
findings of the HMMR expression investigation were examined in
connection with several target genes.
2.6 GEPIA2
Using a common processing pipeline, GEPIA2, a comprehensive online tool
([93]http://gepia2.cancer-pku.cn/), analyzes the RNA sequencing
expression data on 9,736 tumors and 8,587 normal samples from the TCGA
and GTEx projects. Customizable services include differential
expression analysis of tumors vs. normal, profiling based on the
pathological stages or cancer kinds, studies of patient survival,
identification of comparable genes, correlation analysis, and
dimensionality reduction analysis. RNA-seq datasets generated by a
standard procedure from the UCSC Xena project were utilized by GEPIA2
([94]Tang et al., 2019). Using the GEPIA2 portal, the HMMR expression
in pan carcinomas was evaluated. We also generated a heat map showing
the patterns of HMMR expression in several TCGA tumors.
2.7 Analyzing networks involving active genes and the interactions of disease
target genes within the pathways
Utilizing Cytoscape 3.8.0 ([95]http://cytoscape.org/.ver.3.8.0), we
established an intricate network of interactions involving specific
genes associated with the cell cycle, a target gene implicated in
malignancy, and a gene pathway. Cytoscape ([96]Kohl et al., 2011), an
open-source software platform, brings about the visualization of
biological pathways and interactions of molecular networks. Moreover,
these networks can be integrated with annotations, diverse data types,
and gene expression profiles for comprehensive analysis using
Cytoscape.
2.8 Modelling of proteins and protein–protein docking
Molecular modeling of HMMR was conducted through a threading approach
in AlphaFold, [97]Google Colab. The resultant best model underwent
additional refinement processes, addressing steric clashes and
incorporating hydrogen and missing atoms. This refinement was executed
using the Galaxy server ([98]https://galaxy.seoklab.org/). All the
refined models underwent comprehensive quality analysis using SAVES
v6.0, and their secondary structure information was extracted from the
Ramachandran plot. Subsequent to the quality assessment,
protein–protein docking of the modeled proteins aimed at understanding
the interactions between HMMR, AURKA, and CDK1. This docking procedure
was performed utilizing the HDOCK web server. Post-analysis of
protein–protein interactions was carried out through Maestro
(Schrodinger LLC., United States).
2.9 Molecular docking
The exploration of molecular docking involved the target proteins HMMR
and the ligands rapamycin and Torin2. AutoDock version v 4.2.6
([99]https://autodock.scripps.edu/) was utilized for these
investigations. Each docking analysis consisted of three iterations,
resulting in a total of 50 solutions for each case. The specified
parameters included 2,500,000 evaluations, a population size of 500,
and a maximum of 27,000 generations, while the remaining factors were
maintained at the default values. Following the docking procedure, RMSD
collection maps were generated by re-clustering at tolerance levels of
0.5 Å, 1 Å, and 2 Å. This iterative process aimed to pinpoint the
optimal cluster characterized by a substantial population count and the
lowest score of energy.
2.10 Molecular dynamics simulation
Molecular dynamics (MD) simulations were carried out on complexes
involving HMMR with rapamycin, HMMR with Torin2, HMMR with AURKA, and
HMMR with CDK1 utilizing Desmond 2020.1 by Schrödinger, LLC. The
OPLS-2005 force field and an explicit solvent model incorporating TIP3P
water molecules enclosed within a recurrent boundary salvation box
(10 Å × 10 Å x 10 Å) were employed ([100]Jorgensen et al., 1983). To
balance the charge at 0.15 M, Na + ions were introduced, and the NaCl
solution was added to mimic a biological environment. The system
underwent equilibration using an NVT ensemble for 10 ns to stabilize
the protein–ligand complexes. Subsequently, a brief equilibration and
minimization phase was conducted using an NPT ensemble for 12 ns. The
NPT ensemble utilized the Nose–Hoover chain coupling scheme for
temperature fluctuation with a relaxation time of 1.0 ps, and pressure
was maintained at 1 bar throughout every simulation. Employing a time
step of 2 fs, the Martyna–Tuckerman–Klein chain coupling scheme was
employed for pressure control with a relaxation time of 2 ps. The
method that calculated long-range electrostatic interactions with a
fixed radius for Coulomb interactions set at 9 Å was the particle-mesh
Ewald method. Utilizing a RESPA integrator which had a time step of 2
fs for each trajectory, the final production run extended for 200 ns
each. To evaluate the consistency of the MD simulations, various
parameters, including root-mean-square deviation (RMSD),
root-mean-square fluctuation (RMSF), the radius of gyration (Rg), and
the number of hydrogen bonds, were computed ([101]Martyna et al., 1992;
[102]Martyna et al., 1994; [103]Toukmaji and Board Jr, 1996).
2.11 Binding free energy analysis
The determination of binding free energies for ligand–protein complexes
involved employing the generalized Born surface area (MM-GBSA) approach
in conjunction with molecular mechanics. The Prime MM-GBSA binding free
energy was computed using the thermal mmgbsa.py Python script, which
accessed the simulation trajectory for the last 50 frames with a 1-step
sampling size. The calculation of Prime MM-GBSA binding free energy (in
kcal/mol) adhered to the principle of additivity. This entailed summing
up individual energy components, encompassing covalent, hydrogen bond,
van der Waals, columbic, self-contact, lipophilic, and solvation
energies of both the protein and the ligand. The formula employed for
computing ΔG[bind] is outlined as follows:
[MATH:
∆Gbi
nd=ΔG
MM+ΔGSolv
−ΔGSA. :MATH]
Here,
* -
[MATH: ∆ :MATH]
G[bind] depicts the binding free energy.
* -
[MATH: ∆ :MATH]
G[MM] depicts the difference between the free energies of
ligand–protein complexes and the total energies of the protein and
ligand in the isolated form.
* -
[MATH: ∆ :MATH]
G[Solv] represents the difference in the GSA solvation energies of
the ligand–receptor complex and the sum of the solvation energies
of the receptor and the ligand in the unbound state.
* -
[MATH: ∆ :MATH]
G[SA] signifies the difference in the surface area energies of the
protein and the ligand.
3 Results
3.1 Overexpression and upregulation of HMMR across various cancer types
The analysis of HMMR expression patterns was examined through The
Cancer Genome Atlas (TCGA) datasets utilizing TIMER 2.0 and GEPIA2
analyses. The research revealed increased HMMR expression across
numerous malignancies, as depicted in the heat map ([104]Figures 1,
[105]2). Furthermore, the level of mRNA for HMMR was scrutinized using
the UALCAN database by comparing TCGA tumor samples with their
corresponding normal samples ([106]Figures 3, [107]4). It was revealed
that HMMR was highly upregulated in all carcinomas, including ESCA,
COAD, CESC, and READ, followed by BRCA and STAD.
FIGURE 1.
[108]FIGURE 1
[109]Open in a new tab
Differential expression of HMMR within carcinoma and adjacent normal
tissues across all TCGA tumors from the TIMER 2.0 database. The box
plots demonstrate that HMMR is highly upregulated in several
malignancies. The statistical significance was computed by the Wilcoxon
test and is annotated by the number of stars (*p-value<0.05;
**p-value<0.01; ***p-value<0.001).
FIGURE 2.
[110]FIGURE 2
[111]Open in a new tab
Differential gene expression of HMMR between cancerous and adjacent
normal tissues across all TCGA tumors from the GEPIA 2 database. The
heat map demonstrates that HMMR is highly overexpressed in several
malignancies.
FIGURE 3.
[112]FIGURE 3
[113]Open in a new tab
Analysis of mRNA level expression; (A) Expression of HMMR across TCGA
tumours (B) Expression of HMMR across TCGA cancers (with tumour and
normal samples).
FIGURE 4.
[114]FIGURE 4
[115]Open in a new tab
Analysis of HMMR on the basis of (A) sample types, (B) individual
cancer stages, (C) histologic subtypes, (D) breast cancer sub-classes,
(E) major subclasses, and (F) patient age. The bar graphs depict the
elevated levels of HMMR at all levels in breast cancer.
The investigation unveiled an upregulation and overexpression of HMMR
across almost all tumor samples, including various breast cancer
stages, such as individual cancer stages, histological subtypes, breast
cancer sub-classes, TNBC subtypes, and those based on patient age
([116]Figure 4). All these databases represented the significant
overexpression of HMMR in breast cancer at all stages and levels, as
compared to normal samples and its involvement and importance as a
therapeutic target.
3.2 Protein–protein interaction through STRING and Cytoscape
As per the PPI diagram depicted in [117]Figure 5, examining the shared
target genes reveals a network consisting of 11 nodes and 47 edges.
Furthermore, to construct a PPI network based on the amalgamated
network targets, PPI data from the STRING platform were central
proteins within the network. These findings imply that the chosen
components have a strong affinity, making them promising gene targets
for addressing breast carcinoma. The data were incorporated into the
Cytoscape application. Notably, target genes such as AURKA, CDK1 and
TPX2 exhibited an elevated frequency of protein interactions suggesting
their potential role as central proteins within the network. These
findings imply that the chosen components have a strong affinity,
making them promising gene targets for addressing breast carcinoma.
FIGURE 5.
FIGURE 5
[118]Open in a new tab
Outcomes arising from the interplay among interconnected networks of
shared target genes. The protein–protein interaction (PPI) network
representing these common target genes is presented, with nodes
representing the target genes and their relationships visualized
through edges. This network effectively communicates the connections
between the target genes. The nodes are color-coded; cyan and purple
denote confirmed interactions; green, blue, and purple indicate
anticipated interactions; and yellow, sky blue, and light green
represent other interactions. The combination of node colors and their
spatial arrangement provides insights into the three-dimensional
configuration of the target genes.
3.3 Pathway enrichment analysis and Gene Ontology
An exploration of Gene Ontology for the shared target genes indicated a
primary focus on biological processes related to mitotic sister
chromatid segregation and mitotic spindle organization. In terms of
cellular components, the emphasis was on spindle and microtubule
cytoskeleton. Molecular functions included DNA replication origin
binding, single-stranded DNA helicase activity, and microtubule binding
([119]Figure 6). These findings underscore a distinct involvement of
HMMR in the cell cycle, encompassing multiple pathways and intricate
interactions among them.
FIGURE 6.
[120]FIGURE 6
[121]Open in a new tab
(A). Analysis of the processes of utmost significance based on the
numbers of associated target genes and the outcomes of the Gene
Ontology (GO) categories. (B). Heat map representation of the
involvement of HMMR with other genes in biological, cellular, and
molecular processes.
3.4 Gene correlation analysis
The data obtained from PPI interactions ([122]Figure 5) and Gene
Ontology ([123]Figure 6) revealed certain potential target genes, such
as AURKA, TPX2, and CDK1, which could be analyzed to check the
correlation between HMMR and the above-mentioned genes. The heat map
and Pearson’s pairwise correlation plot obtained from bc-GenExMiner
indicated a high correlation between HMMR and the target genes
([124]Figure 7). [125]Figure 8 depicts the gene correlation analysis
through a linear regression graph between HMMR and AURKA, TPX2, and
CDK1 across breast cancer obtained from TIMER 2.0. The graphs show the
highest correlation between HMMR and AURKA, followed by TPX2 and CDK1.
FIGURE 7.
[126]FIGURE 7
[127]Open in a new tab
Analysis of correlation between HMMR and AURKA, TPX2, and CDK1 through
plots obtained from bc-GeneExMiner.
FIGURE 8.
[128]FIGURE 8
[129]Open in a new tab
Gene correlation analysis through a linear regression graph between
HMMR and AURKA, TPX2, and CDK1 across breast cancer where n = 1,100 (n:
no. of patients) obtained from the TIMER 2.0. graphs show the highest
correlation between HMMR and AURKA, followed by TPX2 and CDK1.
3.5 Modeling of protein and validation
Modeled proteins using AlphaFold2 are displayed in [130]Figure 9.
Protein HMMR exhibited two longitudinally arranged α-helices, the
C-terminal domain exhibited a small turn, and the rest conformed to
extended loops at both N and C-terminals ([131]Figure 9). The residual
positions in the formation of the secondary structure were confirmed
using Ramachandran’s plot ([132]Figure 9). Residues are the most
favored region, which was 88.9%, the additional allowed region was
6.25%, and the generously allowed region was 5.2%. No residues were
seen at the disallowed region. Therefore, it could be suggested that
the predicted model is well-validated and considered for further
studies.
FIGURE 9.
[133]FIGURE 9
[134]Open in a new tab
Modeled HHMR protein is in ghostly white cartoon representation;
central panel, Ramachandran’s plot; right panel, 2D plot of the helix
domains.
Molecular docking studies were carried out to decipher the binding
aspects of target HMMR with ligands such as rapamycin and Torin2. The
images of molecular surfaces, docked complexes, and two-dimensional and
three-dimensional interactive plots for targeting HMMR with ligands
such as rapamycin and Torin2 are shown in [135]Figures 10A and B. HMMR
and rapamycin showed a considerable binding affinity of ΔG
−5.4 kcal/mol. The residue Glu317 at the binding cavity was involved in
conventional hydrogen bonding, while other residues of the cavity are
involved in weak van der Waals interactions with rapamycin ([136]Figure
10A). HMMR with the ligand Torin2 exhibited a binding energy of ΔG
−5.9 kcal/mol. Here, the residue Glu328 was involved in conventional
hydrogen bonding, while Phe325, Lys322, and Leu321 were involved in
alkyl and pi–alkyl interactions. Residue Leu321 was also involved in
the pi–sigma interaction ([137]Figure 10B). No other significant
interactions are observed. Protein–protein docking between HMMR + AURKA
and HMMR + CDK1 is shown in [138]Figure 11. The binding energy score
calculated from the HDOCK server for HMMR + AURKA is −738.7, forming
169 non-bonded interactions ([139]Figure 11), a couple of salt bridges,
and five hydrogen bonds. Meanwhile, HMMR + CDK1 exhibited significant
binding and a much lower dock score with the lowest energy of −515.9.
Overall, 164 non-bonded interactions took place where two hydrogen
bonds and a couple of salt bridges were formed ([140]Figure 11).
Therefore, from PPI docking, it could be suggested that HMMR had a
higher interaction with AURKA compared to CDK1.
FIGURE 10.
[141]FIGURE 10
[142]Open in a new tab
(A) Modeled HMMR docked with rapamycin. (B) Modeled HMMR docked with
Torin2 exhibiting the frequency of populations at the 2.0 tolerance
level. Surface view of proteins are exhibiting a deep core of binding
pocket accommodating the ligands, 2D interaction plot of ligands
binding pocket in the respective proteins.
FIGURE 11.
[143]FIGURE 11
[144]Open in a new tab
Structures of protein–protein docking best pose between (A) HMMR +
AURKA and (B) HMMR + CDK1. The middle panel shows the interacting
residue numbers where chain A is for HMMR and explains the number of
residual interactions between HMMR and respective proteins, and the
left panel shows the interacting residues.
3.6 Molecular dynamics simulation
Molecular dynamics and simulation was carried out to assess the
stability and convergence of HMMR in the presence of the drugs
rapamycin and Torin2. The 200-ns simulations revealed consistent
conformations, as evidenced by the comparison of root-mean-square
deviation values. Specifically, the Cα-backbone RMSD of HMMR in complex
with Torin2 displayed a deviation of 2.8 Å, while the protein that was
bound to rapamycin exhibited a similar deviation of 2.8 Å (refer to
[145]Figure 12A). Importantly, all RMSD values remained below the
acceptable threshold of 3 Å. The observed stability in the RMSD plots
throughout the simulation indicates robust convergence and the
maintenance of stable conformations. Hence, it can be proposed that
pharmaceuticals attached to HMMR exhibit considerable stability within
the complex due to the heightened affinity of the ligand. Similarly,
HMMR, when bound to the AURKA protein, demonstrated an RMSD value of
2.91 Å, while with the Cdk1-bound protein, the RMSD was 3.0 Å (see
[146]Figure 12B). The root-mean-square fluctuation (RMSF) plot
displayed minor peaks in fluctuation for the HMMR protein with Torin2,
with the exception of notable spikes at residues 150–155, possibly
indicating increased flexibility in these residues ([147]Figure 12C).
In the case of HMMR with rapamycin, fluctuations occurred at positions
30–45 and 150–175 (refer to [148]Figure 12C). The HMMR bound to AURKA
protein exhibited negligible fluctuations, suggesting a rigid protein
conformation during ligand binding ([149]Figure 12D). Conversely, the
Cdk1-bound protein displayed residual fluctuations at residues 180–230
([150]Figure 12D). Most residues exhibited low fluctuations throughout
the entire 200-ns simulation ([151]Figures 12C and D), signifying
stable amino acid conformations during the simulation period. Thus, the
root-mean-square fluctuation plot suggests that the protein structure
remains rigid during simulation of the conformations that are
ligand-bound.
FIGURE 12.
[152]FIGURE 12
[153]Open in a new tab
MD simulation analysis of 200-ns trajectories of (A) Cα backbone RMSD
of HMMR + Torin2 (red), HMMR + rapamycin (black); (B) Cα backbone RMSD
of HMMR + AURKA (red) and HMMR + CDK1 (black); (C) Cα backbone of HMMR
+ Torin2 (red) and HMMR + rapamycin (black); (D) Cα backbone RMSF of
HMMR + AURKA (red) and HMMR + CDK1 (black). MD simulation analysis of
200-ns trajectories of (E) Cα backbone radius of gyration (Rg) of HMMR
+ Torin2 (red) and HMMR + rapamycin (black). (F) radius of gyration
(Rg) of the Cα backbone of HMMR + AURKA (red) and HMMR + CDK1 (black).
(G) Formation of hydrogen bonds in HMMR + Torin2 (red) and HMMR +
rapamycin (black). (H) Formation of hydrogen bonds in HMMR + AURKA
(red) and HMMR + CDK1 (black).
The radius of gyration (Rg) serves as an indicator of the compactness
of proteins. In this investigation, the Cα-backbone of HMMR bound to
Torin2 consistently displayed an Rg value ranging from 14.8 to 15.0 Å
([154]Figure 12E). In contrast, a distinct pattern was observed for the
rapamycin-bound protein, with Rg values ranging from 14.8 to 14.7 Å
([155]Figure 12E). The Rg values for HMMR bound to AURKA remained
stable, ranging from 14.8 to 14.9 Å, and for HMMR bound to Cdk1, they
fluctuated between 14.7 and 14.76 Å ([156]Figure 12F). A significantly
stable gyration (Rg) suggests a greatly compact orientation of the
protein in the ligand-bound state. A significant interaction and
stability of the complex are indicated by the number of hydrogen bonds
between the protein and the ligand. The hydrogen bond analysis between
HMMR and Torin2 revealed a notable count of two bonds, and with
rapamycin, three hydrogen bonds were observed ([157]Figure 13G).
Similarly, between AURKA and HMMR, two hydrogen bonds were identified,
and with CDK1, there were two hydrogen bonds observed, persisting
throughout the entire 200-ns simulation period ([158]Figure 12H).
FIGURE 13.
[159]FIGURE 13
[160]Open in a new tab
Free energy landscape (FEL) is depicted concerning principal components
(PCs) for conformational scrutiny, where the left panel displays a 2D
FEL along with clusters of frames. The structures at the midpoint are
presented with a time scale, where the right panel showcases the well
of global minima in a 3D representation. (A) HMMR + Torin2, (B) HMMR +
rapamycin, (C) HMMR + AURKA, and (D) HMMR + CDK1.
3.7 Molecular mechanics generalized Born surface area
The binding free energy and additional contributing energy in the form
of MM-GBSA were determined for each, HMMR + rapamycin and HMMR +
Torin2. The binding free energy and additional contributing energy in
the form of molecular mechanics generalized Born surface area were
calculated for each of the two molecules, HMMR + rapamycin and HMMR +
Torin2, using the molecular dynamics simulation trajectory. The
findings ([161]Table 1) indicated that ΔG[bind]Coulomb, ΔG[bind]vdW,
and ΔG[bind]Lipo contributed the most to ΔG[bind] in the stability of
the simulated complexes, but ΔG[bind]Covalent and ΔG[bind]SolvGB
contributed to the instability of the corresponding complexes. The
binding free energies of the HMMR + rapamycin and HMMR + Torin2
complexes were considerably higher. These findings confirmed the
strength of HMMR-containing rapamycin and Torin2 molecules. They also
showed that these molecules could form stable protein–ligand complexes
and were successful in binding to the specified protein.
TABLE 1.
Binding free energy components for the HMMR + Torin2 and HMMR +
rapamycin complex calculated by molecular mechanics generalized Born
surface area.
Energy (kcal/mol) HMMR + Torin2 HMMR + rapamycin
ΔG[bind] −28.53 ± 4.1 −26.15 ± 1.13
ΔG[bind]Lipo −19.83 ± 2.3 −13.43 ± 1.6
ΔG[bind]vdW −12.68 ± 2.17 −14.160 ± 3.0
ΔG[bind]Coulomb −2.14 ± 1.01 −6.22 ± 0.99
ΔG[bind]H[bond] −0.06 ± 0.01 −0.62 ± 0.16
ΔG[bind]SolvGB 13.65 ± 2.27 21.2 ± 1.7
ΔG[bind]Covalent 0.85 ± 0.5 2.66 ± 1.12
[162]Open in a new tab
3.7.1 Principal component analysis and free energy landscape
Principal component analysis (PCA) is conducted on the MD simulation
trajectories of proteins to interpret the randomly selected,
statistically significant conformations (overall movement) of the atoms
within the amino acid residues sampled throughout the trajectory, as
illustrated in the figure. The movement of internal coordinates in
three-dimensional space over a temporal duration of 100 ns was captured
in a covariance matrix. The directional motion of each trajectory is
explained as orthogonal collections or eigenvectors. During the
trajectory, PCA highlights the statistically notable conformations. It
allows for the identification of principal movements within the
trajectory and the essential motions necessary for conformational
alterations. The analysis involved examining two separate aspects,
namely, the distance between Cα-atoms (PC1) and the dihedral angles Φ
and Ψ (PC2) on 2D planes, as illustrated in [163]Figure 13. This
assessment allowed us to identify the primary movements occurring
throughout the trajectory, including crucial reaction coordinates.
Within the HMMR + Torin2 complex, several local minima were observed,
each containing a significant number of frames. The clusters of local
minima highlight significant transition barriers when comparing the
principal components (see [164]Figure 13A). In the case of HMMR +
rapamycin, four to five clusters of local minima were observed. The
presence of the drug-bound state suppresses protein activity and raises
energy barriers. Frames within each specific cluster of local minima
exhibit correlated motion individually, and the eigenvalues are
recorded along both PC1 and PC2 axes (refer to [165]Figure 13B). In the
case of the AURKA protein complexed with HMMR protein, a solitary
prominent cluster was observed within the local minima, indicating a
deep energy well with fewer transition barriers, owing to reduced
protein inhibition (see [166]Figure 13C). For HMMR + CDK1, three
distinct clusters of local minima islands were identified (see
[167]Figure 13D). These discrete clusters of energy minima suggest
significant energy barriers due to drug inhibition, resulting in the
protein undergoing conformational changes from its native folding
pattern.
3.7.2 Analysis of the secondary structure of proteins
Upon analyzing the secondary structural elements (SSEs) crucial for the
protein’s overall stability, it was noted that HMMR + Torin2 exhibited
an approximate average of 41% SSE ([168]Figure 14A), predominantly
comprising helices rather than strands, whereas HMMR + rapamycin
displayed an average of about 40% SSE ([169]Figure 14B) in both its apo
and ligand-bound states. During the 100-ns simulation, minimal
alterations were noted in the snapshots of both the Torin2- and
rapamycin-bound states of the HMMR protein structure. Conversely, in
the HMMR + AURKA and HMMR + CDK1 configurations, the average secondary
structural elements (SSEs) exhibited patterns akin to those observed in
the Torin2-bound state, with the percentage of SSE remaining at 41%
([170]Figures 14C and D). Throughout the simulation, the secondary
structure of the HMMR protein remained relatively constant when bound
to the hits, indicating the hits’ stability in the complex with both
Torin2- and rapamycin-bound HMMR. Hence, the MD simulation indicates
the stability of the complexes.
FIGURE 14.
FIGURE 14
[171]Open in a new tab
Secondary structure element percentage of (A) HMMR + Torin2, (B) HMMR +
rapamycin, (C) HMMR + AURKA, and (D) HMMR + CDK1.
4 Discussion
As breast cancer is an issue of global health concern, it is very
important for researchers to find potential avenues to target it and
find potential therapeutic strategies. Therefore, in this regard, the
versatile functions of HMMR, both extracellularly and intracellularly,
underscore its significance in cancer biology ([172]Sullivan et al.,
2018). Its extracellular role is as a receptor for hyaluronic acid
(HA)-regulated cell migration, a pivotal aspect in inflammation and
wound healing ([173]Aya and Stern, 2014). Intracellularly, HMMR’s
association with the cell cycle, spindle assembly, and microtubules
further solidifies its impact on cell growth and division ([174]Nguyen
et al., 2017). The fine-tuned regulation of HMMR in healthy tissues
contrasts with its upregulation in proliferative tissues, particularly
evident in various cancer types. This dysregulation contributes to
invasiveness and metastasis, aligning with unfavorable prognosis in
cancers such as colorectal, breast, and prostate carcinomas. The
overexpression of HMMR in the mouse mammary epithelium influences the
tumor microenvironment and cancer cell phenotype, leading to an
increase in the genesis of Brca1-mutant tumors ([175]Mateo et al.,
2022).
The consistent overexpression of HMMR across various cancers,
especially in breast cancer, underscores its potential as a significant
biomarker. The UALCAN, TIMER 2.0, and GEPIA2 analyses reveal a
substantial upregulation of HMMR in tumor samples, indicating its role
in disease progression. The robustness of these findings across
multiple databases enhances the credibility of HMMR as a clinically
relevant marker. The AlphaFold2-based molecular modeling of HMMR
provides a reliable structural framework for understanding its
interactions. The subsequent docking studies with rapamycin and Torin2
offer insights into potential therapeutic interventions. The binding
affinities and interaction patterns, especially the involvement of key
residues, provide a rational basis for considering these compounds as
inhibitors. The specificity of these interactions, as illustrated in
the 2D and 3D plots, strengthens the argument for their candidacy as
therapeutic agents. The 200-ns MD simulations provide a dynamic view of
the stability of HMMR-ligand complexes and demonstrate the robustness
of the structures, with minimal deviations and fluctuations. This
stability, coupled with the maintenance of hydrogen bonds, underscores
the potential of rapamycin and Torin2 to form enduring complexes with
HMMR. Molecular mechanics generalized Born surface area-based binding
free energy calculations offer quantitative measures of the energetics
involved in HMMR–ligand interactions. Favorable binding energies for
rapamycin and Torin2 suggest strong and stable binding. The breakdown
of energy components highlights the significance of various forces
contributing to the overall stability of the complexes.
Principal component analysis of molecular dynamics simulation
trajectories provides valuable insights into the conformational
dynamics of protein–ligand complexes. By analyzing the movement of
atoms within amino acid residues, PCA identifies statistically
significant conformations and principal movements essential for
understanding protein dynamics ([176]Shukla and Tripathi, 2020). In our
study, PCA revealed distinct conformational landscapes for HMMR in
complexes with different ligands. The HMMR + Torin2 complex exhibited
multiple local minima, indicating significant transition barriers.
Conversely, HMMR + rapamycin displayed fewer clusters of local minima,
suggesting suppressed protein activity and elevated energy barriers due
to the drug-bound state. This implies different modes of interaction
and potential therapeutic implications for each complex. Notably, the
HMMR + AURKA complex showed a prominent energy well with reduced
transition barriers, indicative of a stable protein–ligand interaction.
In contrast, HMMR + CDK1 exhibited multiple clusters of local minima,
suggesting substantial energy barriers and conformational changes,
potentially impacting protein function.
Analysis of SSEs further supported the stability of HMMR–ligand
complexes. Both HMMR + Torin2 and HMMR + rapamycin maintained
consistent secondary structure percentages throughout the simulation,
indicating stable interactions between HMMR and the respective ligands.
Similarly, HMMR + AURKA and HMMR + CDK1 configurations exhibited stable
secondary structures, suggesting robust binding of HMMR with these
proteins. Overall, our MD simulations demonstrate the stability of
HMMR–ligand complexes and provide insights into their conformational
dynamics. These findings contribute to our understanding of
HMMR-mediated cellular processes and have implications for the
development of targeted therapies in breast cancer treatment.
The proposed mechanism linking HMMR to the CDK1 complex provides
insights into its potential role in accelerating the initiation of
mitosis. This suggests a potential avenue for genetic modifications
supporting tumor growth through increased CDK1/Cyclin B activity
([177]Fulcher and Sapkota, 2020). The correlation between HMMR and key
genes such as AURKA, TPX2, and CDK1 highlights its involvement in
crucial cellular processes. Importantly, these genes are the key
regulators of the mTOR pathway, which is potentially known to be
deregulated in various cancers. This provides a novel perspective on
HMMR’s contribution to breast cancer progression. Notably, we exposed a
plausible interaction between HMMR and AURKA, resulting in an elevation
of AURKA protein levels. This augmentation subsequently triggered the
mTORC2/AKT pathway, as proved in the case of prostate cancer ([178]Miao
et al., 2023). Therefore, we hypothesized that HMMR could be targeted
through mTOR, which serves as a critical mediator in advancing the
progression of breast cancer. The integration of the bioinformatics
analyses and in silico studies leads to a compelling argument for the
clinical relevance of HMMR and its potential as a therapeutic target.
The proposed strategy of targeting the mTOR pathway with rapamycin and
Torin2, based on the strong correlation with HMMR and associated genes,
introduces a novel and promising avenue for breast cancer treatment.
The proposed link between HMMR and the mTOR pathway signifies the
potential therapeutic significance of mTOR inhibitors in breast cancer
treatment. By disrupting this pathway, not only can we target HMMR but
also modulate its associated genes, presenting a comprehensive approach
to curb tumor growth.
Having said that, despite indications that increased levels of HMMR
expression serve as predictive elements for breast cancer patients, all
the data examined in our study originated from bioinformatics databases
and computational studies. Furthermore, owing to its rare occurrence,
there was a deficiency in publicly available data, potentially leading
to statistical imprecision. To address this issue, additional datasets
containing larger cohorts need to be incorporated. Hence, further
inquiries are warranted to ascertain whether HMMR might be harnessed as
diagnostic indicators or therapeutic targets in breast cancer, and
therefore, it is crucial to acknowledge that further experimental
validations, preclinical studies, and clinical trials are necessary to
translate these findings into practical therapeutic applications. This
study not only contributes to the understanding of HMMR’s role in
cancer but also suggests novel therapeutic strategies. The outcomes
suggest that HMMR may function as a critical oncogenic controller and a
promising target for treating breast cancer.
5 Conclusion and future perspective
In conclusion, our study underscores the pivotal role of HMMR in cancer
progression, providing a foundation for targeted therapeutic
interventions. By unraveling the intricate networks involving HMMR and
its correlation with key cellular processes, we pave the way for
innovative strategies aimed at disrupting cancer growth and improving
patient outcomes. The detailed structural insights and dynamic
behaviors obtained from molecular modeling and dynamics simulations
contribute to the overall understanding of HMMR as a potential
therapeutic target in breast cancer. The findings not only highlight
the clinical implications of HMMR but also propose a targeted
therapeutic approach with specific inhibitors. Further research in this
direction holds the potential to translate these findings into
effective clinical applications for breast cancer and beyond.
Nevertheless, an in-depth investigation and in vitro and in vivo
studies for the validation of the link between HMMR and the mTOR
pathway are significant to elucidate the precise function and mechanism
of HMMR in breast cancer progression, which we will try to elucidate in
our further studies.
Acknowledgments