Next Article in Journal
Dynamics of Neutralizing Antibody and T-Cell Responses to SARS-CoV-2 and Variants of Concern after Primary Immunization with CoronaVac and Booster with BNT162b2 or ChAdOx1 in Health Care Workers
Next Article in Special Issue
A Vaccine Construction against COVID-19-Associated Mucormycosis Contrived with Immunoinformatics-Based Scavenging of Potential Mucoralean Epitopes
Previous Article in Journal
Factors Predicting Rubella Vaccination and Antibody in Pregnant Women in Japan: A Report from Pregnant Women Health Initiative
Previous Article in Special Issue
Annotation of Potential Vaccine Targets and Design of a Multi-Epitope Subunit Vaccine against Yersinia pestis through Reverse Vaccinology and Validation through an Agent-Based Modeling Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Immunoinformatics and Biophysics Approaches to Design a Novel Multi-Epitopes Vaccine Design against Staphylococcus auricularis

by
Roba Attar
1,
Eid A. Alatawi
2,
Faris F. Aba Alkhayl
3,4,
Khloud Nawaf Alharbi
3,
Khaled S. Allemailem
3 and
Ahmad Almatroudi
3,*
1
Department of Biology, Faculty of Science, University of Jeddah, Jeddah 21959, Saudi Arabia
2
Department of Medical Laboratory Technology, Faculty of Applied Medical Sciences, University of Tabuk, Tabuk 71491, Saudi Arabia
3
Department of Medical Laboratories, College of Applied Medical Sciences, Qassim University, Buraydah 51452, Saudi Arabia
4
Department of Pharmaceutical Chemistry and Pharmacognosy, College of Dentistry and Pharmacy, Buraydah Colleges, Buraydah 51418, Saudi Arabia
*
Author to whom correspondence should be addressed.
Vaccines 2022, 10(5), 637; https://doi.org/10.3390/vaccines10050637
Submission received: 12 March 2022 / Revised: 6 April 2022 / Accepted: 14 April 2022 / Published: 19 April 2022
(This article belongs to the Special Issue Computer and Technology Supported Development of Vaccines)

Abstract

:
Due to the misuse of antibiotics in our daily lives, antimicrobial resistance (AMR) has become a major health problem. Penicillin, the first antibiotic, was used in the 1930s and led to the emergence of AMR. Due to alterations in the microbe’s genome and the evolution of new resistance mechanisms, antibiotics are losing efficacy against microbes. There are high rates of mortality and morbidity due to antibiotic resistance, so addressing this major health issue requires new approaches. Staphylococcus auricularis is a Gram-positive cocci and is capable of causing opportunistic infections and sepsis. S. auricularis is resistant to several antibiotics and does not currently have a licensed vaccine. In this study, we used bacterial pan-genome analysis (BPGA) to study S. auricularis pan-genome and applied a reverse immunology approach to prioritize vaccine targets against S. auricularis. A total of 15,444 core proteins were identified by BPGA analysis, which were then used to identify good vaccine candidates considering potential vaccine filters. Two vaccine candidates were evaluated for epitope prediction including the superoxide dismutase and gamma-glutamyl transferase protein. The epitope prediction phase involved the prediction of a variety of B-Cell and T-cell epitopes, and the epitopes that met certain criteria, such as antigenicity, immunogenicity, non-allergenicity, and non-toxicity were chosen. A multi-epitopes vaccine construct was then constructed from all the predicted epitopes, and a cholera toxin B-subunit adjuvant was also added to increase vaccine antigenicity. Three-dimensional models of the vaccine were used for downward analyses. Using the best-modeled structure, binding potency was tested with MHC-I, MHC-II and TLR-4 immune cells receptors, proving that the vaccine binds strongly with the receptors. Further, molecular dynamics simulations interpreted strong intermolecular binding between the vaccine and receptors and confirmed the vaccine epitopes exposed to the host immune system. The results support that the vaccine candidate may be capable of eliciting a protective immune response against S. auricularis and may be a promising candidate for experimental in vitro and in vivo studies.

1. Introduction

Antibiotics are medications that are used to completely eradicated or decrease bacteria growth. Antibiotics cannot treat infections caused by viruses [1]. The first antibiotic was penicillin which was discovered in 1928 by Alexander Fleming [2]. The overuse of antibiotics is a major concern as this causes many bacterial infections resistant to antibacterial drugs causing antimicrobial resistance (AMR) [3]. There is a high rate of morbidity and mortality for those pathogens which show resistance to antibiotics. Around 70,000 total deaths are caused due to AMR each year. Due to the over or misuse of antibiotics, bacteria become resistant to antibiotics by making changes in their genetic makeup [4]. To lower the burden of antibiotic resistance, it is essential to make a vaccine that is safe to use and can stop the antibiotic resistance evolution [5]. Resistance against vaccines is very rare and thus could be a more attractive option to manage bacterial pathogens [6].
Vaccines are biological formulations used to stimulate and train the host immune system against infectious agents by producing antibodies and cellular immunity [6]. The first vaccine was developed against smallpox by Edward Jenner by using the virus named cowpox [7]. Different approaches are used in the development of good vaccines and are divided into many types [8]. Then, Louis Pasteur developed a vaccine by using the bacillus in its weakened form to treat anthrax [9]. After Pasteur’s efforts, an extensive number of experiments were conducted in the field for developing vaccines against several diseases. Sabin and stalk also developed a good vaccine against poliovirus using the rules of Pasteur vaccinology [10]. A vaccine named Bacillus Calmette Guerin (BCG) was developed against Mycobacterium tuberculosis using Pasteur vaccinology [11]. However, there are some limitations of Pasteur vaccinology as it does not apply to the pathogens which constantly show mutations in antigens present on surfaces or those pathogens whose in vitro culturing is not possible [12]. Furthermore, developing a conventional vaccine is costly and requires a lot of time [13]. Due to these hurdles, a reduction in the development of a culture-based vaccine is seen, and new techniques are used to develop a vaccine [14]. In the last few years, numerous advancements in vaccinology have been reported. In modern vaccinology called reverse vaccinology (RV), we determine the antigenic surface proteins from a database of the genome [15,16]. In RV we identify good vaccine candidates by applying different vaccine filters [17,18]. Through RV we developed a vaccine named meningococcal serogroup B (4CMenB) [19]. In this technique, vaccine targets are identified from genomic data of the pathogen and predicted B-cell and T-cell epitopes are used in constructing a good vaccine candidate [19,20].
Staphylococcus auricularis, an uncommon coagulase-negative Staphylococcus (CoNS), is commonly considered culture contamination. S. auricularis is a Gram-positive species of the genus Staphylococcus composed of two or four cocci cells. S. auricularis is a popular colonizer of the ear canal. It is not frequently found to be harmful and has only very infrequently been linked to skin and soft tissue diseases [21]. Because it often resides on human skin, it may be capable of causing opportunistic infectious diseases or sepsis, although this is quite rare. Although staphylococci with no coagulase activity may be a normal part of the skin’s flora, these bacteria can cause infection of the skin and soft tissues [22]. It has been observed that S. auricularis causes skin and soft tissue infections [23]. Patients who are immunosuppressed or elderly are more likely to contract these infections, and they are often susceptible to antibiotic treatment. Ear canals are a common site of S. auricularis colonization [21]. It is one of the many hundreds of species of bacteria that constitute normal human flora and is rarely a cause of disease. S. auricularis causes sepsis in newborn babies [23]. S. auricularis is cultivated in cultures of abscesses and paronychia [24]. Thus, coagulase-negative S. auricularis should not always be recognized as pollutants or typical flora, but rather as causal pathogens [21,25]. Several vaccine design efforts have been conducted against pathogenic species of the Staphylococci genus [25,26]. However, the S. auricularis genome has not been explored so far for vaccine targets. As no licensed vaccine is available and the bacterium is evolving new antibiotic resistant mechanisms via horizontal gene transfer, therefore, it needs to be managed by developing a safe and effective vaccine for the pathogen [27,28,29,30].

2. Research Methodology

Figure 1 depicts the methods used to build a multi-epitope peptide vaccine for S. auricularis.

2.1. Bacterial Pan-Genomics, Subtractive Proteomics, and Reverse Vaccinology

Proteomic data of S. auricularis was retrieved from NCBI’s genome database [31]. After that, the proteomes were screened for prospective vaccine candidates through bacterial pan-genomics [32,33], reverse vaccinology [34], and subtractive proteomics [35].

2.2. Pre-Selection Stage

A bacterial pan-genome analysis (BPGA) technique was used to identify the core proteome of S. auricularis [36]. In BPGA analysis, only sequenced genomes were considered and similarity-based clustering was performed at 50%. In BLASTp [31] against human proteome (tax id: 190), the identity of the sequences must be lower than 30%. The non-homologous proteins are ideal candidates for vaccine design as they do not generate auto-immune responses [26,37]. Similarly, to identify pathogenic essential proteins, BLASTp against the DEG database [38] was run, and the proteins that have a sequence similarity score of ≥30% were considered essential [39].

2.3. CD-Hit Analysis

Redundant proteins are duplicate copies of genes and, therefore, are not regarded as effective immunological targets [40]. Non-redundant proteins are the best choices for vaccines [41]. All of the pathogen’s non-redundant proteins were anticipated with an online CD-HIT server [42] with a sequence similarity threshold of 50% and all other input values left at defaults settings. When it comes to comparing and clustering protein sequences, CD-HIT has become a popular and commonly utilized server [43].

2.4. Subcellular Localization Phase

Subcellular organization of the essential and non-redundant proteome was examined through PSORTb 3.0 [44]. Proteins located at the cell surface of infectious agents play an important role in vaccine design because they interact with hosts and aid in the pathogen’s infectious cycle [45].

2.5. Vaccine Candidate’s Prioritization Phase

The pathogenic secretome and exoproteome were further filtered in this stage to find those associated with pathogen pathogenesis and disease progression. The screened proteins present at the pathogen surface or secretory were BLASTp against the virulent factor database (VFDB) [46] for the identification of virulent proteins using sequence specificity of less than 30% and a bit score of greater than 100 [12].

2.6. Prediction of Immune Cell Epitopes

Bepipred Linear Epitope Prediction 2.0 [47] from the Immune Epitope Data source (IEDB) [48] with a value of 0.5 was used to anticipate the first linear B-cell epitopes for virulent proteins. The B-cell epitopes were then utilized to locate T-cell antigenic determinants using IEDB T-cell antigenic determinants analysis packages to find subsequences that interact with MHC (I and II) alleles [49]. The MHC-I interacting epitopes were predicted using IEDB recommended 2020.09 (NetMHCpan EL 4.4) method considering MHC source species as human and reference HLA allele reference data set while MHC-II epitopes were determined using IEDB recommended 2.22 method. The MHC-I presents epitopes to cytotoxic T-cells while MHC-II alleles represent epitopes to helper T-cells [50,51].

2.7. MHCPred 2.0 Analysis

MHCPred 2.0 [52] evaluation was performed to determine the binding affinities of screened B-cell derived T-cell antigenic epitopes, and those with an IC50 value less than 100 nm for DRB*0101 were considered [53].

2.8. Antigenicity, Allergenicity, and Adhesion Probability Prediction

Using VaxiJen 2.0 [54] with a targeted organism set as bacteria and a threshold of >0.4, the antigenicity of predicted epitopes was determined [55]. The accuracy of antigenic targets prediction by VaxiJen varies from organism to organism. For instance, in the case of bacterial antigens, the VaxiJen accuracy is 82%, specificity is 72% and sensitivity is 91% [54]. Allertop 2.0 [56] was used to assess the allergenicity of epitopes [57]. The adhesion properties of the antigenic epitopes were also investigated [58]. As a prospective vaccine target, adhesive proteins let bacteria bind and adhere to host cells, which is essential for infectious pathogenesis. SPAAN was used to estimate the adhesion properties of the antigenic proteins with a baseline range of 0.5 [59].

2.9. Multi-Epitopes Peptide Designing

Peptide vaccines have weak immunogenicity that can be solved by integrating immunodominant epitopes to build a multi-epitope peptide vaccine and using suitable adjuvants [60]. GPGPG linkers were utilized to assemble the filtered epitopes into a multi-epitope peptide [61]. The 3D structure of the design was simulated using the 3Dpro of the SCRATCH protein predictor [62]. Galaxy Loop and Galaxy Refine of Galaxy Web were used to simulate the structure’s loops and optimize the structure further [63,64]. Design 2.0 was used to conduct the introduction of the disulfide bonds in the proposed vaccine structure [49]. Design 2.0 uses the Disulfide by Design algorithm which is developed for recognizing protein folds and native geometry [65].

2.10. Codon Optimization

The vaccine’s design sequence was reverse translated and optimized for codon usage following the E. coli codon usage pattern. The vaccine sequence was reverse translated using Java Codon Adaptation Tool (JCat) service [66,67]. The percent GC concentration and the codon adaptation score (CAI) were used to evaluate the expression rate of the cloned sequence. Ideally, the CAI should be 1. It is recommended that the GC content should be around 30–70 percent because of the high efficiency of transcription and translation.

2.11. Docking and Refinement

In this phase, the recombinant vaccine construct was docked with suitable immune receptors. Using a “blind docking” technique, the vaccine constructs natural binding to TLR4 carrying PDB ID: 4G8A, MHC-I (PDB ID: 1I1Y), and MHC-II (PDB ID: 1GKO) was predicted. PATCHDOCK, an online network that permits the docking of molecules based on shape complementarity principles, was used to execute molecular docking [68,69]. The complex type was set to “default” and the clustering RMSD was set to “4.0” [70]. Fast Interaction Refinement in Molecular Docking (FireDock) [71] was utilized to fine-tune the docked complexes instantly. Using Fire dock’s rescoring and refining tools, protein–protein docking solutions could be quickly improved. A low-global-binding-energy complex was chosen for each case by UCSF Chimera [72], which also used intermolecular interaction and binding confirmation selection [73].

2.12. Molecular Dynamics (MD) Simulation Assay

In silico technologies, such as molecular dynamics, modeling can be used to study the dynamic behavior of vaccine-immune receptors [74]. Complexes were chosen for the molecular dynamic simulation process based on their global energy values [73]. The research was carried out using AMBER20 simulation software [75] on a 250 ns timescale. The AMBER SANDER module is used to complete the system setup, preprocessing, and production phases. Solvating the compounds into a TIP3P solvation box with padding of 12 Angstroms was the final step. Afterward, the complexes were roasted to 300 K and cooled for 1 ns. As part of this process, each trajectory file was saved at the rate of 10 ns per second. To restrict hydrogen bonds, the SHAKE method [76] was used, while Langevin dynamics was employed to regulate the temperature [77,78]. CPPTRAJ module [79] was used for trajectories analysis.

2.13. Free Energy of Immune Receptors and Vaccine Design

For docked complexes, the MMPBSA.py tool [80] provided in AMBER20 was used to determine the binding free energies. In total, 100 frames were selected from the trajectory of a complete simulation and examined for binding energy. A key goal was to identify the differences in free energy between a solvated and an unsolvated state of complexes.

3. Results

3.1. Genomes Retrieval of S. auricularis

For developing a vaccine based on multi-epitopes, sequenced genomes of S. auricularis were retrieved from the genome database of NCBI. Here, we downloaded eleven genome sequences with both complete and incomplete genomes of S. auricularis. The strain size of these pathogens ranges from 2.22 Mb to 2.42 Mb while the GC content is from 37.10 to 37.40. A strain’s type, genome size, and percentage of GC content are presented in Table 1.

3.2. Bacterial Pan-Genome Analysis

Using bacterial pan-genome analysis, we derived the core genome and accessory genome for downward steps. While the pan-genome includes the sequences of all strains, the core genome contains the sequences that are present in all strains [36]. The accessory genome includes the sequences that occur in only a few strains but are not found in all strains [36]. Unique genes are found only in one strain and are strain-specific, also called singletons. Adjoining or dispensable genes are part of the accessory proteome [36]. Those proteins that are conserved across strains are found in the core genome. Figure 2 shows the genome size for each strain, while Figure 3 presents a pan phylogeny tree of S. auricularis.

3.3. CD-HIT Analysis and Proteins Subcellular Localization

The core genome of the pathogen is displayed in Figure 4 and is composed of 1394 non-redundant proteins and 14,050 redundant proteins while the total proteome count is 15,444. Attempts were made to remove the redundant proteins since they are duplicates, and thus, not suitable candidates for vaccine development [81]. Surface proteins, as well as proteins in periplasm, extracellular and outer membrane, are easily recognized by the immune system of the host [82]. As shown in Figure 4, a total of 12 proteins have been identified at the pathogen surface.

3.4. VFDB Analysis

According to the methodology described in the methodology section, two virulent proteins were identified in the selected 12 extracellular/surface proteins as shown in Table 2. Proteins from virulent organisms serve as ideal vaccine targets since they can stimulate immune pathways and result in more effective, safe immune responses [83,84].

3.5. B-Cell Epitopes Prediction

Two proteins were chosen for epitope prediction after passing through all other essential filters. Thereafter, we will use the IEDB server to predict B-cell epitopes and T-cell epitopes. The first step was to predict B-cell epitopes. For gamma-glutamyl transferase, the total predicted B-cell epitopes were eight and for superoxide dismutase four B-cell epitopes were predicted (Table 3).

3.6. MHC-I and MHC-II Epitopes Prediction

The process of predicting T-cell epitopes includes MHC-I and MHC-II binding. These are listed in Table 4. We used the following MHC-I alleles; HLA-A*01:01, HLA-A*01:01, HLA-A*02:01, HLA-A*02:01, HLA-A*02:03, LA-A*02:03, HLA-A*02:06, HLA-A*02:06, HLA-A*03:01, HLA-A*03:01, HLA-A*11:01, HLA-A*11:01, HLA-A*23:01, HLA-A*23:01, HLA-A*24:02, HLA-A*24:02, HLA-A*26:01, HLA-A*26:01, HLA-A*30:01, HLA-A*30:01, HLA-A*30:02, HLA-A*30:02, HLA-A*31:01, HLA-A*31:01, HLA-A*32:01, HLA-A*32:01, HLA-A*33:01, HLA-A*33:01, HLA-A*68:01, HLA-A*68:01, HLA-A*68:02, HLA-A*68:02, HLA-B*07:02, HLA-B*07:02, HLA-B*08:01, HLA-B*08:01, HLA-B*15:01, HLA-B*15:01, HLA-B*35:01, HLA-B*35:01, HLA-B*40:01, HLA-B*40:01, HLA-B*44:02, HLA-B*44:02, HLA-B*44:03, HLA-B*44:03, HLA-B*51:01, HLA-B*51:01, HLA-B*53:01, HLA-B*53:01, HLA-B*57:01, HLA-B*57:01, HLA-B*58:01, HLA-B*58:01) and MHC-II alleles; HLA-DRB1*01:01, HLA-DRB1*03: *04:01, HLA-DRB101,HLA DRB1*04:05, HLADRB1*07:01, HLA,DQA1*03:01/DQB1*03:02,HLADQA1*03:01/DQB1*03:02,HLADQA1*01:02/DQB1*06:02,HLADPA1*02:01/DPB1*01:01,HLADPA1*01:03/DPB1*04:01,HLADPA1*03:01/DPB1*04:02,HLADPA1*02:01/DPB1*05:01,HLADPA1*02:01/DPB1*14:01. The predicted T-cell epitopes are given in Table 4.

3.7. Epitope Prioritization Phase

In order to prioritize those epitopes that can be used in a multi-epitope vaccine, various filters such as water-solubility, antigenicity, MHCPred, allergenicity, and toxicity were applied to the selected epitopes.

3.8. MHCPred Analysis

MHCPred was used to analyze epitope binding affinity for DRB*0101. Only those epitopes with IC50 values less than 100 nm were selected given the prevalence of DRB*0101 in 95% of the population [32]. Table 5 shows those epitopes with IC50 values below 100 nm.

3.9. Allergenicity and Antigenicity

Several non-allergic and antigenic epitopes were taken into consideration to stimulate the strongest immune responses. The epitopes that are non-allergic and antigenic are listed in Table 5.

3.10. Analysis of Solubility and Toxicity

Epitopes with high solubility were identified using InvivoGen and only non-toxic epitopes were selected using Toxin-Pred [85]. In Table 5, non-allergic, water-soluble, non-toxic, and antigenic epitopes were listed. Figure 5 shows the fourteen shortlisted epitopes and how they will be combined into a multi-epitope vaccine.

3.11. Multi-Epitopes Vaccine Designing

The epitopes of the multi-epitope were linked through linkers to allow efficient separation of the epitopes to overcome epitopes’ weak immunogenicity [86]. As well as adding adjuvant molecules to the multi-epitope’s peptide, the immunogenic and antigenic properties of the vaccine were further enhanced [87,88]. A potent interferon stimulator, cholera toxin B-subunit was used as an adjuvant. The schematic representation of the vaccine constructs for multiple epitopes can be found in Figure 6.

3.12. Vaccine Structure Modeling

To investigate the interactions between vaccine constructs and immune receptors, as well as the exposed nature of vaccine epitopes, a three-dimensional model of the vaccine construct was developed. At the time, no suitable template existed for vaccine structure modeling, so Ab initio modeling was performed. A 3D model of the vaccine is shown in Figure 7.

3.13. Loop Modeling and Refinement

Keeping the structure stable following loops which contain residues; TYR33-ILE38, LYS64-PRO74, GLY139-GLY143, GLY167-GLN173, GLN187-ASP192, ASP207-PRO213, GLU232-PRO236, PRO253-GLU262, GLY280-LEU285, LEU300-SER304, PRO311-LYS321, GLU50-LYS55, GLU100-ASN111, GLN152-ALA159, TYR174-ILE179, ILE193-PRO199, GLY214-SER220, GLU237-GLY242, GLY266-GLY270, PRO286-LEU292, ILE305-GLY310, ARG56-PHE63, GLU160-TYR166, SER180-GLY186, GLY200-VAL206, GLY224-GLY228, ASN243-GLY252, PRO271-TYR275, GLU293-ASP299 to obtain the most refined structure, they were modeled into secondary structure elements.

3.14. Disulfide Engineering

Disulfide engineering was applied to the vaccine in order to improve the intermolecular bonding and structural stability of the vaccine. Moreover, it assures that the weaker components of the vaccine are resistant to cellular degradation and impart configurational stability to the vaccine [89]. Cys residues were transformed to cysteine only in groups with a greater energy level (>0 kcal/mol). Cysteine replaces the amino acid residues listed in Table 6, and the cysteine linkages can be seen in Figure 8 as a yellow stick.

3.15. Optimizing Codon Sequences

Following the E. coli expression system, the reversed translation of the vaccine sequence into a DNA sequence was carried out and then optimized for codon usage. The vaccine has a GC level of 57% and a CAI percentage of 0.92%. Both values indicate a high-expressed sequence [90].

3.16. Analysis of Molecular Docking

In order to generate a good immune response, vaccines need to interact effectively with receptors. Through blind docking, we analyze host–receptor interactions with the vaccine construct. Based on the Top 20 docked solutions of vaccines with TLR-4, MHC-I, and MHC-II, are presented in Table 7, Table 8 and Table 9, respectively.

3.17. Docked Complexes Refinement

A further refinement was performed on docked complexes to remove false-positive results and select those with the lowest binding energy. Vaccines that bind best with immune receptors will have the lowest binding energy. For MHC-I, solution number 4 was selected due to its low global energy of 4.17 KJ·m−1. The global binding energy for solution number 9 was −5.24 KJ·m−1 in MHC II. In this case of TLR-4, solution 3 has the lowest global energy of −10.12 KJ·m−1 as compared to any other solution. Docked solutions that have been recorded by FireDock are shown in Table 10, Table 11 and Table 12, respectively.

3.18. Docked Conformation of Vaccine with Immune Receptors

To evaluate docked configuration of vaccine with immunological receptors, such as MHC-I, MHC-II, and TLR-4, the optimal docked complex for every receptor was visualized, as illustrated in Figure 9, Figure 10 and Figure 11. It was discovered that the vaccines had a strong binding affinity for the receptors, allowing the vaccine’s epitopes to be recognized and processed by the body’s immune system. The vaccine’s epitopes can trigger powerful immunological mechanisms and elicit significant and protective immune reactions are additionally implied by this information.

3.19. Interactions of Vaccine to Immune Receptors

Determining the nature and frequency of interactions between the vaccine and receptors is vital, as they determine the intensity of vaccine–receptor interactions. Between the vaccination and receptors, various forms of interactions were identified, including hydrophilic, hydrophobic, salt bridges, and disulfide bonds. They all play a crucial role in the vaccine’s capacity to maintain its docked conformation with the immunological receptors. To involve the vaccine components, these interactions necessitate the presence of many receptor residues. Table 13 shows these residues.

3.20. Molecular Dynamic Simulation

The molecular dynamics simulation of all atoms examined the dynamic behavior of chosen docked complexes. Based on the carbon alpha atoms, we examined the simulation trajectories using the root-mean-square-deviation (RMSD), root-mean-square-fluctuation (RMSF), and hydrogen bonding. To assess if the vaccination epitopes are expressed to the host immune cells and to better comprehend the dynamic binding stability of the vaccine to receptors, this investigation was essential. The RMSD plot remained stable throughout the study, and no significant structural differences were noted. Due to the multiple loops in the systems, a few minor structural abnormalities were observed. According to Figure 12A, the RMSD plot fluctuates around 4–8 Å during a whole simulation run. In addition, the results of the RMSF showed that the key receptor residues retained stability, with just a few high flexibilities because of many loops in the presence of the vaccine molecules. A large percentage of the system’s residues are below 8 Å, indicating that they are more stable than average (Figure 12B). It was found that the systems were compact, and the secondary structures had a tight shape. These findings are comparable to those of RMSD and RMSF and suggest that the system is stable. H-bonds are formed when an electronegative atom of hydrogen bonds with an electronegative atom of another particle. In the case of electronegative acceptors and donor H-bonds, these bonds are formed. The VMD plugin was used to count and identify the H-bonds that originated in the simulation process, as shown in Figure 12C. While the cut-off distance was shown to be 3 Å for all those H-bonds, a pattern of clustering was shown between TLR4, MHC 1 and MHC 11 receptors, and the peptide vaccine construct model. The complex, however, proves to be strong and has a great affinity for binding to the peptide construct and its corresponding immune receptor.

3.21. Estimation of Binding Free Energy

The MM-GBSA and MM-PBSA techniques were used for computing the binding free energy of docked complexes. The fast speed and high efficiency of these two techniques make them highly recognized and reasonable strategies. These estimations were utilized to confirm the binding stability of docked complexes. In MM-GBSA, the total binding free energy of the vaccine-TLR-4 complex is −413.43 kcal/mol, the vaccine-MHC-I complex is −311.01 kcal/mol and for vaccine-MHC-II, the vaccine complex is −270.46 kcal/mol as presented in Table 14. In MM-PBSA the net binding energy is 4–14.23 kcal/mol, −314.55 kcal/mol and −271.82 kcal/mol for vaccine-TLR-4 complex, vaccine-MHC-I complex, and vaccine-MHC-II complex, respectively. Electrostatic and van der Waals energy reveals significant contributions in complexation.

4. Discussion

Traditional vaccines made by inactivating or live-attenuating the infectious agent are effective in inducing immunological responses [9]. However, these vaccines could be allergenic and poisonous because of toxic and potent allergic components. While recombinant vaccines (subunit and conjugates) were developed for targeting a particular protein or toxin, the existence of non-antigenic or allergic factors could result in severe toxicity and vaccine failure [91,92]. Bioinformatics resources are being used to develop multi-epitope vaccines employing a reverse vaccinology strategy [93,94]. Immunoinformatics techniques are being used to reduce the time and expense of vaccine design. The multi-epitopes vaccines are designed with antigenicity, non-toxicity, and non-allergenic properties [95].
In the current research, we used various computational methods to determine epitopes from cell wall adherence proteins of S. auricularis. The retrieval of proteomic data for S. auricularis was carried out from NCBI’s genome database. The prospective vaccine candidates were prioritized through procedures mentioned in the literature. The physicochemical properties were evaluated using web ProtParm software. The transmembrane helices in the selected epitopes were determined using HMMTOP 2.0. The targeted proteins may be proved as efficient vaccine candidates against a variety of S. auricularis strains because they promote cell attachment, nutrition acquisition, and escape from host immune responses. The proteins also cleared the antigenicity, allergenicity, toxicity, and immunogenicity checks that are needed for a successful vaccine design. The antigenicity was evaluated using VaxiJen 2.0, Allertop 2.0 was utilized to determine allergenicity. The adhesive properties of selected epitopes were evaluated by SPAAN. The vaccine epitopes were joined together using particular GPGPG linkers to enhance stability and immunogenicity. Linkers are essential for the functional and structural behavior of fusion proteins and the depiction of specific epitopes in the entire design of the vaccine.
Due to the low proportion of epitopes involved, subunit vaccines have low immunogenicity; however, this can be enhanced with an adjuvant. The beta subunit of cholera toxin was employed as an adjuvant to improve immunogenicity in current immunoinformatics research to develop a new multi-epitope vaccine for bacterial infections caused by S. auricularis [96]. The vaccines designed through multi-epitopes are generally inadequately immunogenic and need adjuvant coupling. Therefore, vaccine design after adjuvant coupling was examined for antigenic and allergenic effects and found to be antigenic and non-allergic. Homology modeling was applied to assess the 3D structure of the vaccine design, which was then accompanied by molecular docking and MD-simulations to evaluate the stability of the association with its particular TLR-4, MHC-1, and MHC-II immune receptors. A well-known TLR-4 is an immune receptor that triggers reactions against a diverse array of ligands and can activate inherent immunologic reactions by causing the formation of pro-inflammatory mediators in response to challenges with S. auricularis [97].
The binding affinities of screened B-cell generated T-cell antigenic factors were determined using MHCPred 2.0. The RMSD and RMSF charts revealed that the vaccine design had a substantial binding affinity for the immune receptor, with minimal deviation. The determination of binding free energy of the TLR4 and multi-epitope vaccine design was carried out using the MMPBSA.py tool. Protein folding and stability depend on disulfide bonding. Furthermore, structural disulfide engineering reduces the range of potential conformations for a particular protein, leading to lowered entropy and enhanced thermostability. The disulfide bond was introduced on the designed vaccine structure using Design 2.0. Despite multi-epitope vaccines being extremely selective to induce a targeted immune response and minimize allergenic responses, in certain situations, the peptide vaccine expressing the modeled epitopes would be slighter immunogenic due to enzymatic degradation in the serum [91].
Purity and stability are two challenges that arise while using peptide-based vaccinations. The most significant problem in peptide vaccine development is incorporating post-translational changes, i.e., glycosylation into recombinant peptides [8]. A computational method for the development and in-silico evaluation of a multi-epitope vaccination against S. auricularis targeting the TLR-4 immune receptor is presented in this research. The multi-epitope vaccine developed in this research must first be examined in animal models and clinical trials to comprehend and investigate the immunological responses of the multi-epitope immunization in humans.

5. Conclusions and Limitations

Several computer-aided vaccine design techniques have been used in this study to create a multi-epitope vaccine against S. auricularis. These techniques include subtractive proteomics, reverse vaccinology, immunoinformatics and several biophysical analyses. Two vaccine targets were used to predict the vaccine epitopes; superoxide dismutase and gamma-glutamyl transferase. The superoxide dismutase has been revealed as a protective antigen against Campylobacter jejuni colonization in chickens. The protein is shown to induce high titer antibodies in chicken serum [98]. Similarly, the gamma-glutamyl transferase is effective in inducing protective immune responses against Helicobacter pylori by promoting effector helper T-cell differentiation into regulatory T-cells [99]. In choosing the targets for vaccine development, several parameters were considered, such as the presence of proteins in the core proteome, present on the cell surface, non-homologous to the host, probiotic bacteria, and feasibility of experimental analysis. In addition, the epitopes that are used in the vaccine are antigenic, not toxic or allergenic, and they bind tightly to B-cell alleles, as well as T-cell alleles. After being simulated, the constructed vaccine showed excellent binding with the different immune receptors. The simulation of host immunity in response to the vaccine revealed a strong primary, secondary, and tertiary immune response. In light of these findings, it was determined that the vaccine was a good candidate to be tested in vivo for immunity. Our findings and data of the study might accelerate the development of vaccines against S. auricularis. We used some pretty tight criteria in selecting subjects for our study, but there are still some shortcomings that need to be addressed in future research. A vaccine’s order of epitopes is not tested for optimal activity. Furthermore, MHC epitopes prediction algorithms are not extensively tested for accuracy.

Author Contributions

Conceptualization, K.S.A. and A.A.; Data curation, R.A., E.A.A., F.F.A.A., K.N.A. and K.S.A.; Formal analysis, R.A., E.A.A., F.F.A.A., K.N.A., K.S.A. and A.A.; Funding acquisition, A.A.; Investigation, R.A., E.A.A., K.N.A. and K.S.A.; Methodology, F.F.A.A., K.N.A. and A.A.; Project administration, A.A.; Resources, A.A.; Software, R.A., E.A.A., F.F.A.A., K.N.A. and K.S.A.; Supervision, A.A.; Validation, K.N.A., K.S.A. and A.A.; Visualization, R.A., E.A.A. and F.F.A.A.; Writing—original draft, R.A., E.A.A. and F.F.A.A.; Writing—review & editing, K.N.A., K.S.A. and A.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available within the article.

Acknowledgments

The researchers would like to thank the Deanship of Scientific Research, Qassim University for funding the publication of this project.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Hutchings, M.; Truman, A.; Wilkinson, B. Antibiotics: Past, present and future. Curr. Opin. Microbiol. 2019, 51, 72–80. [Google Scholar] [CrossRef] [PubMed]
  2. Bennett, J.W.; Chung, K.-T. Alexander Fleming and the Discovery of Penicillin; New Orleans, LA, USA, 2001; Available online: https://www.acs.org/content/acs/en/education/whatischemistry/landmarks/flemingpenicillin.html#alexander-fleming-penicillin (accessed on 11 March 2022).
  3. Hasan, T.H.; Al-Harmoosh, R.A. Mechanisms of antibiotics resistance in bacteria. Sys. Rev. Pharm. 2020, 11, 817–823. [Google Scholar]
  4. O’Neill, J. Tackling Drug-Resistant Infections Globally: Final Report and Recommendations; Wellcome Trust: London, UK, 2016. [Google Scholar]
  5. Adu-Bobie, J.; Capecchi, B.; Serruto, D.; Rappuoli, R.; Pizza, M. Two years into reverse vaccinology. Vaccine 2003, 21, 605–610. [Google Scholar] [CrossRef]
  6. Clem, A.S. Fundamentals of vaccine immunology. J. Glob. Infect. Dis. 2011, 3, 73. [Google Scholar] [CrossRef]
  7. Excler, J.L.; Saville, M.; Berkley, S.; Kim, J.H. Vaccine development for emerging infectious diseases. Nat. Med. 2021, 27, 591–600. [Google Scholar] [CrossRef]
  8. Malonis, R.J.; Lai, J.R.; Vergnolle, O. Peptide-Based Vaccines: Current Progress and Future Challenges. Chem. Rev. 2019, 120, 3210–3229. [Google Scholar] [CrossRef] [Green Version]
  9. Lombard, M.; Pastoret, P.P.; Moulin, A.M. A brief history of vaccines and vaccination. OIE Rev. Sci. Tech. 2007, 26, 29–48. [Google Scholar] [CrossRef] [Green Version]
  10. Kavey, R.-E.W.; Kavey, A.B. Viral Pandemics: From Smallpox to COVID-19; Routledge: London, UK, 2020; Available online: https://www.taylorfrancis.com/books/mono/10.4324/9781003006800/viral-pandemics-rae-ellen-kavey-allison-kavey (accessed on 11 March 2022).
  11. Okafor, C.N.; Rewane, A.; Momodu, I.I. Bacillus Calmette Guerin; StatePearls: Treasure Island, FL, USA, 2019. Available online: https://medlineplus.gov/druginfo/meds/a682809.html (accessed on 11 March 2022).
  12. Naz, A.; Awan, F.M.; Obaid, A.; Muhammad, S.A.; Paracha, R.Z.; Ahmad, J.; Ali, A. Identification of putative vaccine candidates against Helicobacter pylori exploiting exoproteome and secretome: A reverse vaccinology based approach. Infect. Genet. Evol. 2015, 32, 280–291. [Google Scholar] [CrossRef]
  13. Bidmos, F.A.; Siris, S.; Gladstone, C.A.; Langford, P.R. Bacterial Vaccine Antigen Discovery in the Reverse Vaccinology 2.0 Era: Progress and Challenges. Front. Immunol. 2018, 9, 2315. [Google Scholar] [CrossRef] [Green Version]
  14. Buckland, B.C. The process development challenge for a new vaccine. Nat. Med. 2005, 11, S16. [Google Scholar] [CrossRef]
  15. Sette, A.; Rappuoli, R. Reverse vaccinology: Developing vaccines in the era of genomics. Immunity 2010, 33, 530–541. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Suleman, M.; ul Qamar, M.T.; Rasool, S.; Rasool, A.; Albutti, A.; Alsowayeh, N.; Alwashmi, A.S.S.; Aljasir, M.A.; Ahmad, S.; Hussain, Z. Immunoinformatics and Immunogenetics-Based Design of Immunogenic Peptides Vaccine against the Emerging Tick-Borne Encephalitis Virus (TBEV) and Its Validation through In Silico Cloning and Immune Simulation. Vaccines 2021, 9, 1210. [Google Scholar] [CrossRef]
  17. Ul Qamar, M.T.; Shahid, F.; Aslam, S.; Ashfaq, U.A.; Aslam, S.; Fatima, I.; Fareed, M.M.; Zohaib, A.; Chen, L.-L. Reverse vaccinology assisted designing of multiepitope-based subunit vaccine against SARS-CoV-2. Infect. Dis. Poverty 2020, 9, 1–14. [Google Scholar] [CrossRef] [PubMed]
  18. Alharbi, M.; Alshammari, A.; Alasmari, A.F.; Alharbi, S.; Tahir ul Qamar, M.; Abbasi, S.W.; Shaker, B.; Ahmad, S. Whole Proteome-Based Therapeutic Targets Annotation and Designing of Multi-Epitope-Based Vaccines against the Gram-Negative XDR-Alcaligenes faecalis Bacterium. Vaccines 2022, 10, 462. [Google Scholar] [CrossRef] [PubMed]
  19. Serruto, D.; Bottomley, M.J.; Ram, S.; Giuliani, M.M.; Rappuoli, R. The new multicomponent vaccine against meningococcal serogroup B, 4CMenB: Immunological, functional and structural characterization of the antigens. Vaccine 2012, 30, B87–B97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Alharbi, M.; Alshammari, A.; Alasmari, A.F.; Alharbi, S.M.; Tahir ul Qamar, M.; Ullah, A.; Ahmad, S.; Irfan, M.; Khalil, A.A.K. Designing of a Recombinant Multi-Epitopes Based Vaccine against Enterococcus mundtii Using Bioinformatics and Immunoinformatics Approaches. Int. J. Environ. Res. Public Health 2022, 19, 3729. [Google Scholar] [CrossRef]
  21. Natsis, N.E.; Cohen, P.R. Coagulase-negative staphylococcus skin and soft tissue infections. Am. J. Clin. Dermatol. 2018, 19, 671–677. [Google Scholar] [CrossRef]
  22. Roth, R.R.; James, W.D. Microbial ecology of the skin. Annu. Rev. Microbiol. 1988, 42, 441–464. [Google Scholar] [CrossRef]
  23. Hoffman, D.J.; Brown, G.D.; Lombardo, F.A. Early-onset sepsis with Staphylococcus auricularis in an extremely low-birth weight infant-An uncommon pathogen. J. Perinatol. 2007, 27, 519–520. [Google Scholar] [CrossRef] [Green Version]
  24. Ha, E.T.; Heitner, J.F. Staphylococcus Auricularis Endocarditis: A Rare Cause of Subacute Prosthetic Valve Endocarditis with Severe Aortic Stenosis. Cureus 2021, 13, e12738. [Google Scholar] [CrossRef]
  25. Hajighahramani, N.; Nezafat, N.; Eslami, M.; Negahdaripour, M.; Rahmatabadi, S.S.; Ghasemi, Y. Immunoinformatics analysis and in silico designing of a novel multi-epitope peptide vaccine against Staphylococcus aureus. Infect. Genet. Evol. 2017, 48, 83–94. [Google Scholar] [CrossRef] [PubMed]
  26. Tahir ul Qamar, M.; Ahmad, S.; Fatima, I.; Ahmad, F.; Shahid, F.; Naz, A.; Abbasi, S.W.; Khan, A.; Mirza, M.U.; Ashfaq, U.A.; et al. Designing multi-epitope vaccine against Staphylococcus aureus by employing subtractive proteomics, reverse vaccinology and immuno-informatics approaches. Comput. Biol. Med. 2021, 132, 104389. [Google Scholar] [CrossRef] [PubMed]
  27. Szczuka, E.; Jabłońska, L.; Kaznowski, A. Coagulase-negative staphylococci: Pathogenesis, occurrence of antibiotic resistance genes and in vitro effects of antimicrobial agents on biofilm-growing bacteria. J. Med. Microbiol. 2016, 65, 1405–1413. [Google Scholar] [CrossRef] [PubMed]
  28. Khan, M.M.; Faiz, A.; Ashshi, A.M. Clinically significant Coagulase Negative Staphylococci and their antibiotic resistance pattern in a tertiary care hospital. J. Pak. Med. Assoc. 2014, 64, 1171–1174. [Google Scholar]
  29. Morgenstern, M.; Erichsen, C.; Hackl, S.; Mily, J.; Militz, M.; Friederichs, J.; Hungerer, S.; Bühren, V.; Moriarty, T.F.; Post, V.; et al. Antibiotic resistance of commensal Staphylococcus aureus and coagulase-negative staphylococci in an international cohort of surgeons: A prospective point-prevalence study. PLoS ONE 2016, 11, e0148437. [Google Scholar] [CrossRef]
  30. Koksal, F.; Yasar, H.; Samasti, M. Antibiotic resistance patterns of coagulase-negative staphylococcus strains isolated from blood cultures of septicemic patients in Turkey. Microbiol. Res. 2009, 164, 404–410. [Google Scholar] [CrossRef]
  31. Blast, N. Basic local alignment search tool. Natl. Libr. Med. Natl. Cent. Biotechnol. Inf. 2015. Available online: https://blast.ncbi.nlm.nih.gov/Blast.cgi (accessed on 11 March 2022).
  32. Ali, A.; Naz, A.; Soares, S.C.; Bakhtiar, M.; Tiwari, S.; Hassan, S.S.; Hanan, F.; Ramos, R.; Pereira, U.; Barh, D. Pan-genome analysis of human gastric pathogen H. pylori: Comparative genomics and pathogenomics approaches to identify regions associated with pathogenicity and prediction of potential core therapeutic targets. Biomed Res. Int. 2015, 2015, 139580. [Google Scholar] [CrossRef] [Green Version]
  33. Tahir Ul Qamar, M.; Zhu, X.; Xing, F.; Chen, L.-L. ppsPCP: A plant presence/absence variants scanner and pan-genome construction pipeline. Bioinformatics 2019, 35, 4156–4158. [Google Scholar] [CrossRef]
  34. Enayatkhani, M.; Hasaniazad, M.; Faezi, S.; Gouklani, H.; Davoodian, P.; Ahmadi, N.; Einakian, M.A.; Karmostaji, A.; Ahmadi, K. Reverse vaccinology approach to design a novel multi-epitope vaccine candidate against COVID-19: An in silico study. J. Biomol. Struct. Dyn. 2021, 39, 2857–2872. [Google Scholar] [CrossRef] [Green Version]
  35. Ahmad, S.; Raza, S.; Uddin, R.; Azam, S.S. Comparative subtractive proteomics based ranking for antibiotic targets against the dirtiest superbug: Acinetobacter baumannii. J. Mol. Graph. Model. 2018, 82, 74–92. [Google Scholar] [CrossRef] [PubMed]
  36. Chaudhari, N.M.; Gupta, V.K.; Dutta, C. BPGA-an ultra-fast pan-genome analysis pipeline. Sci. Rep. 2016, 6, 24373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Ismail, S.; Shahid, F.; Khan, A.; Bhatti, S.; Ahmad, S.; Naz, A.; Almatroudi, A.; ul Qamar, M.T. Pan-Vaccinomics Approach Towards a Universal Vaccine Candidate Against WHO Priority Pathogens to Address Growing Global Antibiotic Resistance. Comput. Biol. Med. 2021, 136, 104705. [Google Scholar] [CrossRef] [PubMed]
  38. Zhang, R.; Ou, H.-Y.; Zhang, C.-T. DEG: A database of essential genes. Nucleic Acids Res. 2004, 32, D271–D272. [Google Scholar] [CrossRef]
  39. Azam, S.S.; Shamim, A. An insight into the exploration of druggable genome of Streptococcus gordonii for the identification of novel therapeutic candidates. Genomics 2014, 104, 203–214. [Google Scholar] [CrossRef]
  40. Rehman, A.; Ahmad, S.; Shahid, F.; Albutti, A.; Alwashmi, A.S.S.; Aljasir, M.A.; Alhumeed, N.; Qasim, M.; Ashfaq, U.A.; Tahir ul Qamar, M. Integrated Core Proteomics, Subtractive Proteomics, and Immunoinformatics Investigation to Unveil a Potential Multi-Epitope Vaccine against Schistosomiasis. Vaccines 2021, 9, 658. [Google Scholar] [CrossRef]
  41. Abbas, G.; Zafar, I.; Ahmad, S.; Azam, S.S. Immunoinformatics design of a novel multi-epitope peptide vaccine to combat multi-drug resistant infections caused by Vibrio vulnificus. Eur. J. Pharm. Sci. 2020, 142, 105160. [Google Scholar] [CrossRef]
  42. Fu, L.; Niu, B.; Zhu, Z.; Wu, S.; Li, W. CD-HIT: Accelerated for clustering the next-generation sequencing data. Bioinformatics 2012, 28, 3150–3152. [Google Scholar] [CrossRef]
  43. Li, W.; Godzik, A. Cd-hit: A fast program for clustering and comparing large sets of protein or nucleotide sequences. Bioinformatics 2006, 22, 1658–1659. [Google Scholar] [CrossRef] [Green Version]
  44. Solanki, V.; Tiwari, M.; Tiwari, V. Prioritization of potential vaccine targets using comparative proteomics and designing of the chimeric multi-epitope vaccine against Pseudomonas aeruginosa. Sci. Rep. 2019, 9, 5240. [Google Scholar] [CrossRef]
  45. Barh, D.; Barve, N.; Gupta, K.; Chandra, S.; Jain, N.; Tiwari, S.; Leon-Sicairos, N.; Canizalez-Roman, A.; Rodrigues dos Santos, A.; Hassan, S.S. Exoproteome and secretome derived broad spectrum novel drug and vaccine candidates in Vibrio cholerae targeted by Piper betel derived compounds. PLoS ONE 2013, 8, e52773. [Google Scholar] [CrossRef] [PubMed]
  46. Liu, B.; Zheng, D.; Jin, Q.; Chen, L.; Yang, J. VFDB 2019: A comparative pathogenomic platform with an interactive web interface. Nucleic Acids Res. 2019, 47, D687–D692. [Google Scholar] [CrossRef] [PubMed]
  47. Jespersen, M.C.; Peters, B.; Nielsen, M.; Marcatili, P. BepiPred-2.0: Improving sequence-based B-cell epitope prediction using conformational epitopes. Nucleic Acids Res. 2017, 45, W24–W29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Dhanda, S.K.; Mahajan, S.; Paul, S.; Yan, Z.; Kim, H.; Jespersen, M.C.; Jurtz, V.; Andreatta, M.; Greenbaum, J.A.; Marcatili, P. IEDB-AR: Immune epitope database—Analysis resource in 2019. Nucleic Acids Res. 2019, 47, W502–W506. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Pandey, R.K.; Bhatt, T.K.; Prajapati, V.K. Novel immunoinformatics approaches to design multi-epitope subunit vaccine for malaria by investigating anopheles salivary protein. Sci. Rep. 2018, 8, 1125. [Google Scholar] [CrossRef] [Green Version]
  50. Jones, E.Y.; Fugger, L.; Strominger, J.L.; Siebold, C. MHC class II proteins and disease: A structural perspective. Nat. Rev. Immunol. 2006, 6, 271. [Google Scholar] [CrossRef]
  51. Hewitt, E.W. The MHC class I antigen presentation pathway: Strategies for viral immune evasion. Immunology 2003, 110, 163–169. [Google Scholar] [CrossRef]
  52. Guan, P.; Doytchinova, I.A.; Zygouri, C.; Flower, D.R. MHCPred: A server for quantitative prediction of peptide—MHC binding. Nucleic Acids Res. 2003, 31, 3621–3624. [Google Scholar] [CrossRef] [Green Version]
  53. Vakili, B.; Eslami, M.; Hatam, G.R.; Zare, B.; Erfani, N.; Nezafat, N.; Ghasemi, Y. Immunoinformatics-aided design of a potential multi-epitope peptide vaccine against Leishmania infantum. Int. J. Biol. Macromol. 2018, 120, 1127–1139. [Google Scholar] [CrossRef]
  54. Doytchinova, I.A.; Flower, D.R. VaxiJen: A server for prediction of protective antigens, tumour antigens and subunit vaccines. BMC Bioinform. 2007, 8, 4. [Google Scholar] [CrossRef] [Green Version]
  55. Azim, K.F.; Hasan, M.; Hossain, M.N.; Somana, S.R.; Hoque, S.F.; Bappy, M.N.I.; Chowdhury, A.T.; Lasker, T. Immunoinformatics approaches for designing a novel multi epitope peptide vaccine against human norovirus (Norwalk virus). Infect. Genet. Evol. 2019, 74, 103936. [Google Scholar] [CrossRef] [PubMed]
  56. Dimitrov, I.; Bangov, I.; Flower, D.R.; Doytchinova, I. AllerTOP v. 2—A server for in silico prediction of allergens. J. Mol. Model. 2014, 20, 2278. [Google Scholar] [CrossRef] [PubMed]
  57. Hossain, M.S.; Hossan, M.I.; Mizan, S.; Moin, A.T.; Yasmin, F.; Akash, A.-S.; Powshi, S.N.; Hasan, A.K.R.; Chowdhury, A.S. Immunoinformatics approach to designing a multi-epitope vaccine against Saint Louis Encephalitis Virus. Inform. Med. Unlocked 2021, 22, 100500. [Google Scholar] [CrossRef]
  58. Wizemann, T.M.; Adamou, J.E.; Langermann, S. Adhesins as targets for vaccine development. Emerg. Infect. Dis. 1999, 5, 395. [Google Scholar] [CrossRef] [PubMed]
  59. Sajjad, R.; Ahmad, S.; Azam, S.S. In silico screening of antigenic B-cell derived T-cell epitopes and designing of a multi-epitope peptide vaccine for Acinetobacter nosocomialis. J. Mol. Graph. Model. 2020, 94, 107477. [Google Scholar] [CrossRef]
  60. Nain, Z.; Abdulla, F.; Rahman, M.M.; Karim, M.M.; Khan, M.S.A.; Sayed, S.B.; Mahmud, S.; Rahman, S.M.R.; Sheam, M.M.; Haque, Z. Proteome-wide screening for designing a multi-epitope vaccine against emerging pathogen Elizabethkingia anophelis using immunoinformatic approaches. J. Biomol. Struct. Dyn. 2020, 38, 4850–4867. [Google Scholar] [CrossRef]
  61. Saadi, M.; Karkhah, A.; Nouri, H.R. Development of a multi-epitope peptide vaccine inducing robust T cell responses against brucellosis using immunoinformatics based approaches. Infect. Genet. Evol. 2017, 51, 227–234. [Google Scholar] [CrossRef]
  62. Naz, S.; Ahmad, S.; Walton, S.; Abbasi, S.W. Multi-epitope based vaccine design against Sarcoptes scabiei paramyosin using immunoinformatics approach. J. Mol. Liq. 2020, 319, 114105. [Google Scholar] [CrossRef]
  63. Qamar, M.T.U.; Saba Ismail, S.A.; Mirza, M.U.; Abbasi, S.W.; Ashfaq, U.A.; Chen, L.-L. Development of a Novel Multi-Epitope Vaccine Against Crimean-Congo Hemorrhagic Fever Virus: An Integrated Reverse Vaccinology, Vaccine Informatics and Biophysics Approach. Front. Immunol. 2021, 12, 669812. [Google Scholar] [CrossRef]
  64. Dorosti, H.; Eskandari, S.; Zarei, M.; Nezafat, N.; Ghasemi, Y. Design of a multi-epitope protein vaccine against herpes simplex virus, human papillomavirus and Chlamydia trachomatis as the main causes of sexually transmitted diseases. Infect. Genet. Evol. 2021, 96, 105136. [Google Scholar] [CrossRef]
  65. Craig, D.B.; Dombkowski, A.A. Disulfide by Design 2.0: A web-based tool for disulfide engineering in proteins. BMC Bioinform. 2013, 14, 346. [Google Scholar] [CrossRef] [Green Version]
  66. Hossan, M.I.; Chowdhury, A.S.; Hossain, M.U.; Khan, M.A.; Mahmood, T.B.; Mizan, S. Immunoinformatics aided-design of novel multi-epitope based peptide vaccine against Hendra henipavirus through proteome exploration. Inform. Med. Unlocked 2021, 25, 100678. [Google Scholar] [CrossRef]
  67. Shankar, U.; Jain, N.; Mishra, S.K.; Sk, M.F.; Kar, P.; Kumar, A. Mining of Ebola virus genome for the construction of multi-epitope vaccine to combat its infection. J. Biomol. Struct. Dyn. 2021, 1–17. [Google Scholar] [CrossRef] [PubMed]
  68. Khatoon, N.; Pandey, R.K.; Prajapati, V.K. Exploring Leishmania secretory proteins to design B and T cell multi-epitope subunit vaccine using immunoinformatics approach. Sci. Rep. 2017, 7, 8285. [Google Scholar] [CrossRef]
  69. Kar, P.P.; Srivastava, A. Immuno-informatics analysis to identify novel vaccine candidates and design of a multi-epitope based vaccine candidate against Theileria parasites. Front. Immunol. 2018, 9, 2213. [Google Scholar] [CrossRef]
  70. Mugunthan, S.P.; Harish, M.C. Multi-epitope-based vaccine designed by targeting cytoadherence proteins of Mycoplasma gallisepticum. ACS Omega 2021, 6, 13742–13755. [Google Scholar] [CrossRef] [PubMed]
  71. Mashiach, E.; Schneidman-Duhovny, D.; Andrusier, N.; Nussinov, R.; Wolfson, H.J. FireDock: A web server for fast interaction refinement in molecular docking. Nucleic Acids Res. 2008, 36, W229–W232. [Google Scholar] [CrossRef] [PubMed]
  72. Pettersen, E.F.; Goddard, T.D.; Huang, C.C.; Couch, G.S.; Greenblatt, D.M.; Meng, E.C.; Ferrin, T.E. UCSF Chimera—A visualization system for exploratory research and analysis. J. Comput. Chem. 2004, 25, 1605–1612. [Google Scholar] [CrossRef] [Green Version]
  73. Chukwudozie, O.S.; Gray, C.M.; Fagbayi, T.A.; Chukwuanukwu, R.C.; Oyebanji, V.O.; Bankole, T.T.; Adewole, R.A.; Daniel, E.M. Immuno-informatics design of a multimeric epitope peptide based vaccine targeting SARS-CoV-2 spike glycoprotein. PLoS ONE 2021, 16, e0248061. [Google Scholar] [CrossRef]
  74. Rafi, S.; Yasmin, S.; Uddin, R. A molecular dynamic simulation approach: Development of dengue virus vaccine by affinity improvement techniques. J. Biomol. Struct. Dyn. 2022, 40, 61–76. [Google Scholar] [CrossRef]
  75. Case, D.A.; Belfon, K.; Ben-Shalom, I.; Brozell, S.R.; Cerutti, D.; Cheatham, T.; Cruzeiro, V.W.D.; Darden, T.; Duke, R.E.; Giambasu, G.; et al. USA Amber 2020. 2020. Available online: https://ambermd.org/doc12/Amber20.pdf (accessed on 11 March 2022).
  76. Kräutler, V.; Van Gunsteren, W.F.; Hünenberger, P.H. A fast SHAKE algorithm to solve distance constraint equations for small molecules in molecular dynamics simulations. J. Comput. Chem. 2001, 22, 501–508. [Google Scholar] [CrossRef]
  77. Khan, M.; Khan, S.; Ali, A.; Akbar, H.; Sayaf, A.M.; Khan, A.; Wei, D.-Q. Immunoinformatics approaches to explore Helicobacter Pylori proteome (Virulence Factors) to design B and T cell multi-epitope subunit vaccine. Sci. Rep. 2019, 9, 13321. [Google Scholar] [CrossRef] [PubMed]
  78. Majee, P.; Jain, N.; Kumar, A. Designing of a multi-epitope vaccine candidate against Nipah virus by in silico approach: A putative prophylactic solution for the deadly virus. J. Biomol. Struct. Dyn. 2021, 39, 1461–1480. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Roe, D.R.; Cheatham, T.E., III. PTRAJ and CPPTRAJ: Software for processing and analysis of molecular dynamics trajectory data. J. Chem. Theory Comput. 2013, 9, 3084–3095. [Google Scholar] [CrossRef] [PubMed]
  80. Miller, B.R.; McGee, T.D.; Swails, J.M.; Homeyer, N.; Gohlke, H.; Roitberg, A.E. MMPBSA.py: An efficient program for end-state free energy calculations. J. Chem. Theory Comput. 2012, 8, 3314–3321. [Google Scholar] [CrossRef] [PubMed]
  81. Sanober, G.; Ahmad, S.; Azam, S.S. Identification of plausible drug targets by investigating the druggable genome of MDR Staphylococcus epidermidis. Gene Rep. 2017, 7, 147–153. [Google Scholar] [CrossRef]
  82. Ali, S.; Ali, S.; Javed, S.O.; Shoukat, S.; Ahmad, S.; Ali, S.S.; Hussain, Z.; Waseem, M.; Rizwan, M.; Suleman, M. Proteome wide vaccine targets prioritization and designing of antigenic vaccine candidate to trigger the host immune response against the Mycoplasma genitalium infection. Microb. Pathog. 2021, 152, 104771. [Google Scholar] [CrossRef]
  83. Ahmad, S.; Azam, S.S. A novel approach of virulome based reverse vaccinology for exploring and validating peptide-based vaccine candidates against the most troublesome nosocomial pathogen: Acinetobacter baumannii. J. Mol. Graph. Model. 2018, 83, 1–11. [Google Scholar] [CrossRef]
  84. Fatima, I.; Ahmad, S.; Abbasi, S.W.; Ashfaq, U.A.; Shahid, F.; ul Qamar, M.T.; Rehman, A.; Allemailem, K.S. Designing of a multi-epitopes-based peptide vaccine against rift valley fever virus and its validation through integrated computational approaches. Comput. Biol. Med. 2021, 141, 105151. [Google Scholar] [CrossRef]
  85. Gupta, S.; Kapoor, P.; Chaudhary, K.; Gautam, A.; Kumar, R.; Raghava, G.P.S. Peptide toxicity prediction. In Computational Peptidology; Springer: Berlin/Heidelberg, Germany, 2015; pp. 143–157. [Google Scholar]
  86. Javadi, M.; Oloomi, M.; Bouzari, S. In Silico Design of a Poly-epitope Vaccine for Urinary Tract Infection Based on Conserved Antigens by Modern Vaccinology. Int. J. Pept. Res. Ther. 2021, 27, 909–921. [Google Scholar] [CrossRef]
  87. Ullah, A.; Ahmad, S.; Ismail, S.; Afsheen, Z.; Khurram, M.; Tahir ul Qamar, M.; AlSuhaymi, N.; Alsugoor, M.H.; Allemailem, K.S. Towards A Novel Multi-Epitopes Chimeric Vaccine for Simulating Strong Immune Responses and Protection against Morganella morganii. Int. J. Environ. Res. Public Health 2021, 18, 10961. [Google Scholar] [CrossRef] [PubMed]
  88. Naz, S.; Ahmad, S.; Abbasi, S.W.; Ismail, S.; Waseem, S.; Ul Qamar, M.T.; Almatroudi, A.; Ali, Z. Identification of immunodominant epitopes in allelic variants VK210 and VK247 of Plasmodium Vivax Circumsporozoite immunogen. Infect. Genet. Evol. 2021, 96, 105120. [Google Scholar] [CrossRef]
  89. Zakeri, B.; Fierer, J.O.; Celik, E.; Chittock, E.C.; Schwarz-Linek, U.; Moy, V.T.; Howarth, M. Peptide tag forming a rapid covalent bond to a protein, through engineering a bacterial adhesin. Proc. Natl. Acad. Sci. USA 2012, 109, E690–E697. [Google Scholar] [CrossRef] [Green Version]
  90. Ali, M.; Pandey, R.K.; Khatoon, N.; Narula, A.; Mishra, A.; Prajapati, V.K. Exploring dengue genome to construct a multi-epitope based subunit vaccine by utilizing immunoinformatics approach to battle against dengue infection. Sci. Rep. 2017, 7, 9232. [Google Scholar] [CrossRef] [PubMed]
  91. Li, W.; Joshi, M.D.; Singhania, S.; Ramsey, K.H.; Murthy, A.K. Peptide vaccine: Progress and challenges. Vaccines 2014, 2, 515–536. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Aslam, S.; Ahmad, S.; Noor, F.; Ashfaq, U.A.; Shahid, F.; Rehman, A.; Tahir ul Qamar, M.; Alatawi, E.A.; Alshabrmi, F.M.; Allemailem, K.S. Designing a Multi-Epitope Vaccine against Chlamydia trachomatis by Employing Integrated Core Proteomics, Immuno-Informatics and In Silico Approaches. Biology 2021, 10, 997. [Google Scholar] [CrossRef] [PubMed]
  93. Elliott, S.L.; Suhrbier, A.; Miles, J.J.; Lawrence, G.; Pye, S.J.; Le, T.T.; Rosenstengel, A.; Nguyen, T.; Allworth, A.; Burrows, S.R. Phase I trial of a CD8+ T-cell peptide epitope-based vaccine for infectious mononucleosis. J. Virol. 2008, 82, 1448–1457. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Gul, S.; Ahmad, S.; Ullah, A.; Ismail, S.; Khurram, M.; Tahir ul Qamar, M.; Hakami, A.R.; Alkhathami, A.G.; Alrumaihi, F.; Allemailem, K.S. Designing a Recombinant Vaccine against Providencia rettgeri Using Immunoinformatics Approach. Vaccines 2022, 10, 189. [Google Scholar] [CrossRef]
  95. Greenwood, B. The contribution of vaccination to global health: Past, present and future. Philos. Trans. R. Soc. B Biol. Sci. 2014, 369, 20130433. [Google Scholar] [CrossRef] [Green Version]
  96. Peschel, A.; Otto, M. Phenol-soluble modulins and staphylococcal infection. Nat. Rev. Microbiol. 2013, 11, 667–673. [Google Scholar] [CrossRef]
  97. Mukherjee, S.; Karmakar, S.; Babu, S.P.S. TLR2 and TLR4 mediated host immune responses in major infectious diseases: A review. Braz. J. Infect. Dis. 2016, 20, 193–204. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Chintoan-Uta, C.; Cassady-Cain, R.L.; Al-Haideri, H.; Watson, E.; Kelly, D.J.; Smith, D.G.E.; Sparks, N.H.C.; Kaiser, P.; Stevens, M.P. Superoxide dismutase SodB is a protective antigen against Campylobacter jejuni colonisation in chickens. Vaccine 2015, 33, 6206–6211. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Oertli, M.; Noben, M.; Engler, D.B.; Semper, R.P.; Reuter, S.; Maxeiner, J.; Gerhard, M.; Taube, C.; Müller, A. Helicobacter pylori γ-glutamyl transpeptidase and vacuolating cytotoxin promote gastric persistence and immune tolerance. Proc. Natl. Acad. Sci. USA 2013, 110, 3047–3052. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. A schematic representation of the approach used to identify vaccine candidates in the S. auricularis proteome.
Figure 1. A schematic representation of the approach used to identify vaccine candidates in the S. auricularis proteome.
Vaccines 10 00637 g001
Figure 2. Each strain’s genome size is presented herein. On the x-axis number of genomes is provided while on the y-axis number of gene families in each genome is presented.
Figure 2. Each strain’s genome size is presented herein. On the x-axis number of genomes is provided while on the y-axis number of gene families in each genome is presented.
Vaccines 10 00637 g002
Figure 3. A pan-phylogenetic tree of the genome of S. auricularis.
Figure 3. A pan-phylogenetic tree of the genome of S. auricularis.
Vaccines 10 00637 g003
Figure 4. The total number of core proteins, sub-cellular localized proteins, redundant, non-redundant proteins and virulent proteins. The proteins obtained at each step were screened from core proteome of 11 strains of S. auricularis.
Figure 4. The total number of core proteins, sub-cellular localized proteins, redundant, non-redundant proteins and virulent proteins. The proteins obtained at each step were screened from core proteome of 11 strains of S. auricularis.
Vaccines 10 00637 g004
Figure 5. Fourteen shortlisted epitopes for the construction of a multi-epitopes based vaccine.
Figure 5. Fourteen shortlisted epitopes for the construction of a multi-epitopes based vaccine.
Vaccines 10 00637 g005
Figure 6. The vaccine construct is described in a schematic diagram. In the orange color, we can see the linker (GPGPG) that is used to link selected epitopes. Green represents cholera toxin B-subunit (adjuvant) and purple represents EAAAK (linker).
Figure 6. The vaccine construct is described in a schematic diagram. In the orange color, we can see the linker (GPGPG) that is used to link selected epitopes. Green represents cholera toxin B-subunit (adjuvant) and purple represents EAAAK (linker).
Vaccines 10 00637 g006
Figure 7. The GPGPG linkers (yellow), cholera toxin B subunit (pink), EAAAK linkers (red), and vaccine epitopes (green) are displayed in the 3D structure of the vaccine construct.
Figure 7. The GPGPG linkers (yellow), cholera toxin B subunit (pink), EAAAK linkers (red), and vaccine epitopes (green) are displayed in the 3D structure of the vaccine construct.
Vaccines 10 00637 g007
Figure 8. A mutated and wild vaccine construct. Mutated structures show yellow bands where disulfide bonds have been introduced.
Figure 8. A mutated and wild vaccine construct. Mutated structures show yellow bands where disulfide bonds have been introduced.
Vaccines 10 00637 g008
Figure 9. Docking of the vaccine to the MHC-I molecule. The green color shows the vaccine molecule while blue represents MHC-I.
Figure 9. Docking of the vaccine to the MHC-I molecule. The green color shows the vaccine molecule while blue represents MHC-I.
Vaccines 10 00637 g009
Figure 10. Docking of the vaccine to the MHC-II molecule. The green color shows the vaccine molecule and pink represents MHC-II.
Figure 10. Docking of the vaccine to the MHC-II molecule. The green color shows the vaccine molecule and pink represents MHC-II.
Vaccines 10 00637 g010
Figure 11. Docking of the vaccine to the TLR-4 molecule. The green color shows the vaccine molecule while purple represents TLR-4.
Figure 11. Docking of the vaccine to the TLR-4 molecule. The green color shows the vaccine molecule while purple represents TLR-4.
Vaccines 10 00637 g011
Figure 12. Simulation trajectories analysis via: RMSD (A), RMSF (B), and number of Hydrogen bonds (C).
Figure 12. Simulation trajectories analysis via: RMSD (A), RMSF (B), and number of Hydrogen bonds (C).
Vaccines 10 00637 g012
Table 1. S. auricularis strains with different statistics.
Table 1. S. auricularis strains with different statistics.
Organism NameStrainSize (Mb)GC%
S. auricularisFDAARGOS_8822.2237.40
S. auricularisNCTC121012.2237.40
S. auricularisJCM 24212.2637.40
S. auricularisDSM 206092.2037.30
S. auricularisNCTC 121012.2037.20
S. auricularisDE03812.4237.40
S. auricularisS522.2837.10
S. auricularis1H202.2537.10
S. auricularisSNUC 30342.2837.10
S. auricularisSNUC 9932.3037.10
S. auricularisCHK138−47841.6837.10
Table 2. From the exposed proteins, virulent proteins were identified.
Table 2. From the exposed proteins, virulent proteins were identified.
ProteinsSubcellular LocalizationBit ScoreSequence Identity
core/231/1/Org1_Gene1151
gamma-glutamyl transferase
Extracellular244 bits34%
core/1478/1/Org1_Gene484
superoxide dismutase
Periplasmic179 bits43%
Table 3. B-cell epitopes for selected proteins.
Table 3. B-cell epitopes for selected proteins.
B-Cell EpitopesPeptides
core/231/1/Org1_Gene1151
gamma-glutamyl transferase
PDTYDKNELYENKIAGHTQSASNQ
QNAEPKQYQYKDISSYNFKQ
LYGSDVDRDSPFFDGSRTKREGDVVK
KQLDNKLTKKDFQDYEMT
VNGQDNNSENYLSQE
NEVNQTQEEINPEGIDN
SDPNSPNYGEKHKQ
IDVFKNMGYNVEEKRNDP
core/1478/1/Org1_Gene484
superoxide dismutase
LPNLPYDFDALEPYIDKE
DLENKSIEEIVANLDSVPED
TPNSEEKGTVVDKIKEQWGSL
LKYQNKRPEYIE
Table 4. MHC-I and MHC-II epitope prediction by B-Cell epitopes.
Table 4. MHC-I and MHC-II epitope prediction by B-Cell epitopes.
MHC-IIPercentile ScoreMHC-IPercentile Score
KIAGHTQSASN4.3KIAGHTQSA0.15
AGHTQSASN28
PDTYDKNELYENKI28PDTYDKNELY0.28
DKNELYENKI20
KQYQYKDISSYNFKQ4.8QYQYKDISSY0.01
DISSYNFKQ4.4
NAEPKQYQYKDISS28NAEPKQYQY0.03
KQYQYKDISS7.4
LYGSDVDRDSPFFD7DVDRDSPFF0.41
LYGSDVDRD24
DVDRDSPFFD9.9
SPFFDGSRTKREGDV37FFDGSRTKR0.17
SPFFDGSRTK0.31
GSRTKREGDV13
LTKKDFQDYEM7.2LTKKDFQDY0.2
KKDFQDYEM2.9
KQLDNKLTKKDFQDY33KQLDNKLTK0.05
LTKKDFQDY0.2
GQDNNSENYLSQE45GQDNNSENY0.28
NSENYLSQE7.8
NEVNQTQEEINPE13EVNQTQEEI0.1
NEVNQTQEEI0.79
QTQEEINPE6.3
VNQTQEEINPEGIDN22QEEINPEGI0.63
VNQTQEEINP37
SDPNSPNYGEKHKQ74NSPNYGEKHK3.4
DPNSPNYGEK3.5
SDPNSPNYGE6.8
IDVFKNMGYNV0.82DVFKNGYNV0.34
IDVFKNMGY1.4
VFKNMGYNVEEKRND12MGYNVEEKR0.62
VFKNMGYNV1.3
YNVEEKRND65
PNLPYDFDALEPYI1.1DFDALEPYI1.9
PNLPYDFDAL38
LPYDFDALEPYIDKE LPYDFDALE1.7
DALEPYIDKE14
IEEIVANLDSVPE3.2IEEIVANLD18
VANLDSVPE9.3
DLENKSIEEIVANLD11DLENKSIEEI1.4
SIEEIVANLD17
TVVDKIKEQWGSL22TVVDKIKEQW0.08
KIKEQWGSL0.26
TPNSEEKGTVVDKI43EEKGTVVDKI0.57
TPNSEEKGTV0.89
LKYQNKRPEYI3YQNKRPEYI0.21
Table 5. MHC-Pred results, antigenicity, allergenicity, solubility, and toxin-pred profiling are considered in shortlisting epitopes.
Table 5. MHC-Pred results, antigenicity, allergenicity, solubility, and toxin-pred profiling are considered in shortlisting epitopes.
MHC- PredDRB*0101 IC50 ScoreAntigenicityAllergenicitySolubilityToxin Pred
KNELYENKI7.7680.5035
DISSYNFKQ57.151.5642
NAEPKQYQY2.931.0100
KQYQYKDISS83.370.5079
QYQYKDISS11.020.5677
LYGSDVDRD37.240.9838
GQDNNSENY64.861.5275Non-allergenGood water solubilityNon-toxin
QTQEEINPE6.341.1296
NQTQEEINP22.390.9309
DFDALEPYI9.381.2216
PNLPYDFDA18.541.0231
LPYDFDALE56.360.7775
DLENKSIEE32.431.3709
EEKGTVVDK42.270.8901
Table 6. List of residues replaced by cysteine.
Table 6. List of residues replaced by cysteine.
Sequence NumberAmino AcidSequence NumberAmino AcidChi3EnergySum B-Factors
7GLY34HIS67.212.370
15SER24GLN88.655.880
18TYR23PRO74.334.570
209ASP218ASN122.685.370
216GLN220SER122.934.950
225PRO229GLN113.023.010
238GLY245THR110.795.250
285LEU288ASP106.063.040
301GLU305ILE106.283.680
Table 7. Vaccine solutions docked for the top 20 MHC-I vaccines.
Table 7. Vaccine solutions docked for the top 20 MHC-I vaccines.
Solution NoScoreAreaACETransformation
1201323445.10405.18−0.44 0.33 1.63 99.21 38.67 29.15
2198443938.00311.82−0.92 0.91 0.60 22.26–27.15 9.21
3192822805.20378.13−2.47–0.49 3.11–18.74 12.00 37.06
4192723180.70371.95−0.55–0.04 2.27 92.68 70.24 41.81
5182182781.40283.06−0.95 0.40–3.05 42.02 21.11 3.99
6179862731.90493.30−1.97 0.09 0.80 2.95 47.68 44.62
7179683816.80300.18−3.01 1.42 1.28–14.91–40.50–6.83
8179402845.80427.92−2.81 0.37–2.54 46.23–29.01 36.16
9179363052.6086.01−2.91–0.90 2.92–3.45 14.79 55.77
10179343182.80498.192.21 0.23–1.45 52.65–6.49 5.15
11179042907.60493.90−0.97 0.93 0.88 33.16–26.75 13.49
12176882908.80139.472.77–0.08–0.59 64.71 52.46 56.37
13176742832.70465.570.17–0.26 2.33 14.30–1.73–34.33
14175022380.50477.76−0.85–0.07 0.54 22.31 1.39 66.68
15171682920.20471.122.24 0.17–0.67 28.98 58.30–17.67
16169222557.70389.82−0.39 1.28 2.15 47.56 2.78–27.80
17168663903.90336.25−1.45 0.79 1.08 27.04–33.05 26.94
18168262201.00447.60−0.07–0.17–2.61–7.26 71.44 30.42
19167823966.20232.52−0.83–0.06–1.50 2.64 41.63 19.99
20167763039.60442.93−0.33 0.92–2.63–9.19 69.57 23.36
Table 8. Vaccine solutions docked for the top 20 MHC-II vaccines.
Table 8. Vaccine solutions docked for the top 20 MHC-II vaccines.
Solution NoScoreAreaACETransformation
1205703565.40271.97−2.17 –0.39 –2.54 108.70 52.17 53.60
2203823218.70−25.73–1.94 –0.37 2.71 87.91 105.99 58.18
3191963055.90213.631.68 0.11 1.94 111.16 36.90 –54.35
4189923096.60143.820.14 –0.74 –1.31 39.48 110.23 –5.13
5185922957.80177.571.48 –0.90 1.70 138.56 40.32 –13.49
6184883716.80101.29–1.83 –0.24 2.67 94.90 105.50 59.00
7179422977.00343.89−0.38 –0.06 0.02 117.98 14.99 –19.99
8177802450.90395.910.82 0.11 –1.88 125.52 53.90 –48.68
9177363020.90355.921.52 0.18 1.93 115.95 39.51 –55.51
10176124129.70451.991.30 –0.16 1.77 110.37 47.24 –52.75
11175402462.60404.36−1.18 0.51 –1.31 117.85 80.56 6.15
12173102888.70355.39−3.03 0.48 –1.16 105.35 115.94 –6.39
13172322316.70477.49−2.65 –0.81 –2.79 76.34 64.59 43.98
14171362466.10424.10−2.22 0.14 –0.88 133.03 42.01 49.26
15170962801.00425.081.22 –0.46 1.74 94.04 58.55 –31.78
16170742166.70249.61−3.05 0.31 0.14 88.68 108.92 35.75
17168782952.00489.28−1.91 0.06 2.00 123.27 28.22 11.90
18168322913.60347.82−0.77 –0.12 1.64 149.12 75.77 44.31
19168082452.30270.53−0.36 0.38 –1.74 81.35 107.07 26.55
20167843289.80440.55−1.76 –0.04 1.19 102.77 90.46 58.32
Table 9. Vaccine solutions docked for the top 20 TLR4 vaccines.
Table 9. Vaccine solutions docked for the top 20 TLR4 vaccines.
Solution NoScoreAreaACETransformation
1203783608.40443.053.04 −0.64 1.00 −38.53 51.32 −52.20
2202003636.60420.462.82 0.21 −0.74 −50.55 43.33 −63.93
3197303020.20148.512.05 0.19 2.66 −27.29 20.10 −80.32
4194203354.30328.45−0.55 0.18 −2.68 −51.45 45.12 17.38
5189603190.80253.67−1.65 −0.25 −0.03 −36.03 32.54 −32.10
6189423010.20418.13−0.74 −0.23 1.98 −0.03 −42.39 −0.40
7189083164.90159.142.21 0.07 0.42 −62.64 46.15 −9.60
8186542702.20457.752.89 0.54 −1.86 −6.36 25.96 8.73
9184282381.70407.500.19 0.20 2.02 26.92 22.52 −8.30
10181202812.80185.16−0.76 −0.73 −1.94 −27.68 −21.12 14.81
11179362628.00−81.00−0.16 −0.13 2.43 39.24 44.49 −72.70
12178462217.50249.641.79 −0.07 −1.37 30.39 33.31 −87.83
13178162887.90356.04−0.07 −0.95 2.66 46.97 46.18 −40.96
14177802755.00288.751.14 0.48 1.45 4.04 −19.81 −27.50
15177362691.70452.350.96 1.15 −3.01 −7.89 −5.87 −90.50
16175963641.90295.19−2.03 −0.23 −1.28 0.76 48.92 7.22
17173242388.50215.650.78 −0.70 −2.12 16.08 33.32 −89.40
18170063484.90158.111.30 −0.00 2.52 10.41 6.75 −92.90
19170042233.20413.441.14 0.40 1.64 5.01 −17.91 −26.25
20169722286.50443.840.89 −0.21 −2.41 −72.64 11.88 −109.83
Table 10. MHC-I-vaccine FireDock solutions. Energy is measured in KJ·m−1 for each term listed below.
Table 10. MHC-I-vaccine FireDock solutions. Energy is measured in KJ·m−1 for each term listed below.
RankSolution NumberGlobal EnergyAttractive VdWRepulsive VdWACEHB
144.17−31.1619.8411.47−5.30
256.15−0.680.00−0.750.00
3319.06−22.1522.6919.85−0.33
4852.95−14.4749.9111.17−3.19
510455.54−48.39574.7533.65−7.69
66743.07−36.55965.0116.58−8.90
791611.59−20.942061.10−6.63−5.60
813726.53−70.154711.5612.36−7.89
928261.20−123.3510537.7222.74−27.64
10710978.99−103.1413898.947.02−12.71
Table 11. FireDock solutions for MHC-II vaccine. In each of the terms below, the unit of energy is KJ·m−1.
Table 11. FireDock solutions for MHC-II vaccine. In each of the terms below, the unit of energy is KJ·m−1.
RankSolution NumberGlobal EnergyAttractive VdWRepulsive VdWACEHB
19−5.24−4.693.430.930.00
2819.99−16.0344.333.41−2.16
3487.59−20.1693.0012.23−0.25
43266.04−22.59342.765.64−2.57
55323.56−29.38438.659.04−1.39
66411.02−30.17555.416.83−4.74
71869.72−55.341144.9215.69−6.62
871648.44−43.842070.9819.56−4.58
9103313.72−59.534183.4430.29−4.02
1023450.81−76.674459.199.78−10.96
Table 12. Solutions of TLR-4 vaccine by FireDock. For each term below, KJ·m−1 represents the equivalent amount of energy.
Table 12. Solutions of TLR-4 vaccine by FireDock. For each term below, KJ·m−1 represents the equivalent amount of energy.
RankSolution NumberGlobal EnergyAttractive VdWRepulsive VdWACEHB
13−10.12−30.0623.125.72−1.33
210−3.55−36.2323.6716.26−1.65
3810.03−0.010.000.050.00
4617.42−19.985.6015.01−0.67
5928.73−25.5341.8914.33−3.45
6531.17−27.308.0918.01−3.54
7133.43−21.4721.4022.06−2.42
84183.74−33.69303.962.38−4.28
92508.87−22.98645.3115.89−7.35
1071460.30−73.921947.38−1.23−6.66
Table 13. Interaction between receptor residues and vaccine molecules.
Table 13. Interaction between receptor residues and vaccine molecules.
Vaccine ComplexInteractive Residues
MHC-IASN 42, ALA 59, ASN 131, ASP 276, ASN 218, ASN 110, GLU 104, GLY 182, GLN 187, GLY 240, GLU 307, HIS 34, ILE 138, ILE 234, LYS162, LYS 5, LEU 13, MET 58, PHE 150, PRO 170, PRO 269, TYR 148, TYR 33, THR 49, VAL 206, VAL 71
MHC-IIASP 43, ASP 178, ALA 159, ASP 290, ALA 53, ARG 56, GLY 203, GLY 268, GLU 247, GLN 77, GLY 200, GLN 175, HIS 20, LEU 52, LYS 84, PRO 251, SER 195, SER 81, TYR 223, TYR 166, TYR 287
TLR-4ASP 28, ALA 59, ARG 94, ALA 119, ASP 144, ASN 302, ASP 320, CYS 30, CYS 107, GLY 184, GLY 256, GLY 296, ILE 45, ILE 68, ILU 133, LYS 162, LYS 315, LEU 200, PHE 10, THR 99, TRP 109, TYR 275VAL 71, VAL 206
Table 14. Binding free energies shown in kcal/mol.
Table 14. Binding free energies shown in kcal/mol.
Energy ParameterTLR-4-Vaccine ComplexMHC-I-Vaccine ComplexMHC-II-Vaccine Complex
MM-GBSA
VDWAALS−280.47−214.36−194.85
Electrostatic−172.96−146.39−105.61
Delta G solv40.0049.7430.00
Delta Total−413.43−311.01−270.46
MM-PBSA
VDWAALS−280.47−214.36−194.85
EEL−172.96−146.39−105.61
Delta G solv39.2046.2028.64
Delta Total−414.23−314.55−271.82
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Attar, R.; Alatawi, E.A.; Aba Alkhayl, F.F.; Alharbi, K.N.; Allemailem, K.S.; Almatroudi, A. Immunoinformatics and Biophysics Approaches to Design a Novel Multi-Epitopes Vaccine Design against Staphylococcus auricularis. Vaccines 2022, 10, 637. https://doi.org/10.3390/vaccines10050637

AMA Style

Attar R, Alatawi EA, Aba Alkhayl FF, Alharbi KN, Allemailem KS, Almatroudi A. Immunoinformatics and Biophysics Approaches to Design a Novel Multi-Epitopes Vaccine Design against Staphylococcus auricularis. Vaccines. 2022; 10(5):637. https://doi.org/10.3390/vaccines10050637

Chicago/Turabian Style

Attar, Roba, Eid A. Alatawi, Faris F. Aba Alkhayl, Khloud Nawaf Alharbi, Khaled S. Allemailem, and Ahmad Almatroudi. 2022. "Immunoinformatics and Biophysics Approaches to Design a Novel Multi-Epitopes Vaccine Design against Staphylococcus auricularis" Vaccines 10, no. 5: 637. https://doi.org/10.3390/vaccines10050637

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

Article Metrics

Back to TopTop