Next Article in Journal
Effects of Multifaceted Determinants on Individual Stress: The Mediating Role of Social Capital
Next Article in Special Issue
Computer-Aided Multi-Epitope Vaccine Design against Enterobacter xiangfangensis
Previous Article in Journal
Exploring Hepatocellular Carcinoma Mortality Using Weighted Regression Estimation for the Cohort Effect in Taiwan from 1976 to 2015
Previous Article in Special Issue
COVID-19 Clinical Profiles and Fatality Rates in Hospitalized Patients Reveal Case Aggravation and Selective Co-Infection by Limited Gram-Negative Bacteria
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Vaccinomics to Design a Multi-Epitopes Vaccine for Acinetobacter baumannii

1
Department of Health and Biological Sciences, Abasyn University, Peshawar 25000, Pakistan
2
Department of Medical Biotechnology, College of Applied Medical Sciences, Qassim University, Buraydah 52571, Saudi Arabia
3
Department of Biological Sciences, National University of Medical Sciences, Rawalpindi 46000, Pakistan
4
Institute of Molecular Biology and Biotechnology (IMBB), The University of Lahore, Lahore 54590, Pakistan
5
Department of Pharmacy, Abasyn University, Peshawar 25000, Pakistan
6
Department of Theoretical and Applied Chemistry, South Ural State University, Lenin Prospect 76, 454080 Chelyabinsk, Russia
7
Department of Biomedical Engineering, Chung-Ang University, Seoul 06974, Korea
8
Department of Bioinformatics and Biotechnology, Government College University Faisalabad, Faisalabad 38000, Pakistan
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this study.
Int. J. Environ. Res. Public Health 2022, 19(9), 5568; https://doi.org/10.3390/ijerph19095568
Submission received: 27 February 2022 / Revised: 23 April 2022 / Accepted: 28 April 2022 / Published: 4 May 2022
(This article belongs to the Special Issue Clinical Microbiology and Infectious Diseases)

Abstract

:
Antibiotic resistance (AR) is the result of microbes’ natural evolution to withstand the action of antibiotics used against them. AR is rising to a high level across the globe, and novel resistant strains are emerging and spreading very fast. Acinetobacter baumannii is a multidrug resistant Gram-negative bacteria, responsible for causing severe nosocomial infections that are treated with several broad spectrum antibiotics: carbapenems, β-lactam, aminoglycosides, tetracycline, gentamicin, impanel, piperacillin, and amikacin. The A. baumannii genome is superplastic to acquire new resistant mechanisms and, as there is no vaccine in the development process for this pathogen, the situation is more worrisome. This study was conducted to identify protective antigens from the core genome of the pathogen. Genomic data of fully sequenced strains of A. baumannii were retrieved from the national center for biotechnological information (NCBI) database and subjected to various genomics, immunoinformatics, proteomics, and biophysical analyses to identify potential vaccine antigens against A. baumannii. By doing so, four outer membrane proteins were prioritized: TonB-dependent siderphore receptor, OmpA family protein, type IV pilus biogenesis stability protein, and OprD family outer membrane porin. Immuoinformatics predicted B-cell and T-cell epitopes from all four proteins. The antigenic epitopes were linked to design a multi-epitopes vaccine construct using GPGPG linkers and adjuvant cholera toxin B subunit to boost the immune responses. A 3D model of the vaccine construct was built, loop refined, and considered for extensive error examination. Disulfide engineering was performed for the stability of the vaccine construct. Blind docking of the vaccine was conducted with host MHC-I, MHC-II, and toll-like receptors 4 (TLR-4) molecules. Molecular dynamic simulation was carried out to understand the vaccine-receptors dynamics and binding stability, as well as to evaluate the presentation of epitopes to the host immune system. Binding energies estimation was achieved to understand intermolecular interaction energies and validate docking and simulation studies. The results suggested that the designed vaccine construct has high potential to induce protective host immune responses and can be a good vaccine candidate for experimental in vivo and in vitro studies.

1. Introduction

Antibiotic resistance or AR refers to the resistance ability of bacteria to the action of antibiotics to which they were susceptible before. In the last 25 years, bacteria have developed AR very rapidly, and the situation is alarming now [1]. Resistance is rising to a high level across the globe, and novel resistant strains are emerging and spreading very fast [2]. This process occurs due to the evolution of novel mutations in the bacterial genome or acquisition of resistant genes from the environment. Misuse and overuse of antibiotics also contributed significantly to AR [3]. AR genes are transferred from one bacterium to the other through plasmid, which disseminates the resistant genes within/between the species [4]. Some bacteria are resistant to one type of antibiotic while others are multi-drug resistant (MDR), extensive drug resistant (XDR) [5], and pan drug resistant, showing resistance to multiple, many, and all classes of antibiotics, respectively [6]. The number of infections and deaths that occur due to AR is very high. Every year, more than 50,000 deaths occur in America and Europe only due to Methicillin-resistant Staphylococcus aureus (MRSA) [7]. Tuberculosis is still affecting developing countries and leads to a high number of deaths every year [8]. It is tough to manage AR bacterial infections according to the center for disease and control (CDC) survey [9]. Approximately 2.8 million bacterial infections are reported each year, of which more than 35,000 lead to deaths. This means that every 15 min, 1 person dies due to AR [10]. The situation is equally bad all over the world, so immediate actions are needed to combat AR [11]. According to a report by the World Health Organization (WHO), if the situation is not handled properly AR may lead to 10 million deaths and an economic loss of more than 100 trillion dollars every year [12]. In conclusion, we are heading towards the post-antibiotic era, which warrants the introduction of novel therapeutic strategies to address AR health crises [13].
To overcome the burden of AR, many strategies can be applied. This can be achieved by introducing new methodologies that can help in identifying new drug targets and new classes of antibiotics [14]. Traditional technologies of drug target identification and drug development are very slow and costly [15]. Advanced sequencing technologies allow a deeper understanding of bacterial biology and deciphering metabolic processes [16]. This opens up new avenues for developing drugs to manage super drug resistant species that it was not possible to manage before. One way to stop the evolution of AR is the development of antivirulent compounds [16,17]. These compounds only interfere with the pathogenic and virulent attributes of the pathogen but do not force bacteria to change their genome in response to the action of the compound. This is exemplified by the inactivation of quorum sensing signal molecules (through a process known as QS inhibition or quorum quenching (QQ)) to stop bacterial biofilm formation and multiple virulence factors regulators [18,19]. This can be accomplished in several ways, such as through the development of drugs against QS signal molecules and the enzymatic blockage of QS signal molecules pathways. Quorum sensing inhibitors can be used as an alternative to antibiotics [8].
A vaccine is a biological preparation that provides protection against harmful diseases. Vaccination is the most effective method of preventing infections [20]. Historically, vaccines for managing and preventing the spread of AR bacterial pathogens have been underrated, but it is a well-established method to tackle the AR pathogen [21]. For example, Strep pneumoniae (pneumococcal) conjugate and Hemophilus influenza type B (Hib) vaccines are praised for their ability to provide protection against the aforementioned pathogens and lower the use of antibiotics against them [22]. Considering this, vaccines have the potential to provide a permanent solution to tackle AR bacteria. A. baumannii is a Gram-negative bacteria usually found in hospital environments. It causes a variety of nosocomial infections, such as pneumonia, meningitis, wound infections, and urinary tract infections (UTIs), which are difficult to control and prevent [23]. A. baumannii is treated with several broad-spectrum antibiotics: carbapenems, β-lactam, aminoglycosides, tetracycline, gentamicin, imipenem, piperacillin, and amikacin [10]. Over the last 15 years, this bacteria has been a high risk to human health because it has acquired resistance to most antibiotics and has become a serious health threat [24,25]. No vaccine is currently available against this pathogen, which makes the situation more worrisome [26]. Therefore, efforts are needed to identify protective antigenic peptides from the core genome of the pathogen, which can easily be analyzed in experimental studies. Findings of the study will save time and may lead to cost-effective vaccine development.

2. Research Methodology

The study was initiated by retrieving complete sequenced genomes of A. baumannii from the national center for biotechnological information (NCBI) database. Subtractive proteomic and reverse vaccinology (RV) techniques were applied to prioritize vaccine candidates against A. baumannii. This whole process was completed in three phases. The step wise procedure applied herein is presented in Figure 1.

2.1. Complete Genome Retrieval

The complete sequenced genome of Acinetobacter baumannii (~50 in number) was retrieved from NCBI database. The core and dispensable genomes were identified using bacterial pan-genome analysis [24].

2.2. Pre-Screening Phase

The primary phase was named as a pre-screening phase. During this phase, antigenic and virulence factor molecules were selected. Those proteins were selected that (i) showed sequence conservation among all sequenced genomes [27], (ii) were not similar to the human proteome as they may generate autoimmune responses [28], (iii) were present on the surface of the pathogen because of their strong ability to stimulate the host immune system [29], (iv) were essential and crucial for the survival of the pathogen due to their key cellular functions [30], (v) were part of the core proteome [31], and (vi) were non-redundant [32], because redundant proteins are not part of the core genome and are poor immunological targets [33]. On the other hand, non-redundant proteins are good vaccine candidates because they are present in all strains and have vital functions in cells [34]. The redundancy of proteins was predicted through a server called Cd-Hit [35] by considering a sequence identity threshold of 50%. Then, homology of the proteins was checked against the human proteome through an online tool named BlastP [36]. There are different parameters for identifying the similarity between human and bacterial proteins. Proteins with less than 30% sequence identity, a bit score of 100 or more, and an E-value smaller than 1.0 × 10−4 [37] were regarded as host non-homologous. Another server, PsortB, was used to predict the surface localized proteins [38,39]. Only extracellular, periplasmic, and outer membrane localized proteins were subjected to further analysis, and all the cytoplasmic membrane proteins were discarded [40].

2.3. Vaccine Epitopes Prioritization Phase

During this phase, further filtration of the pathogen secretome and exoproteome was conducted. Virulent analysis was carried out to check the involvement of proteins in the pathogen infection pathway [41]. This selection was done via BlastP against the virulent factor database (VFDB) [42]. Proteins (virulence factors) with 30% identity and a bit score higher than 100 were chosen [43]. The physiochemical properties of the selected virulence factors were checked via Protparam [44]. Only proteins that were easy to use in experimental analysis were selected [45]. Proteins with an instability index of less than 40 were selected, and those with more than 40 were marked unstable. The proteins were then analyzed for their molecular weight; those with a molecular weight of less than 110 kDa were considered as the best vaccine candidates; HMMTOP and THMM servers were then used to examine the number of transmembrane helices [46]. Only proteins with 0 and 1 transmembrane helices were selected for further processing [24,29]. The prediction of antigenicity was performed using Vaxijen2.0 [47]. The adhesive nature of the selected antigenic proteins was predicted using another server called SPAAN with a minimum value of 0.5 [48]. To overcome the chances of inhibiting beneficial bacteria, the sequence alignment of filtered proteins was performed out with probiotic bacterial proteomes [49]. For this, 4 lactobacillus bacterial species (L. rhamnosus (taxid: 47,715), L. casei (taxid: 1582), L. gasseri (taxid: 1596) and L. johnsonii (taxid: 33,959)) and one Bifidobacterium specie, i.e., B. bifidum (taxid: 1681), were selected, and a BlastP search was accomplished [50]. Furthermore, the similarity of proteins was checked against mouse proteome just to extract those proteins that were showing no homology to the mouse proteins. This was performed to decrease the chances of autoimmune reactions and reduce false positive results during trails testing of the vaccine candidate [51]. After passing through these filters, the selected proteins were predicted for B-cell and T-cell epitopes [52,53,54]. Linear B cell prediction was performed first using the IEDB server [54]. T-cell prediction was performed on the same IEDB server using B-cell epitopes. Binding to both MHC-I and MHC-II alleles was predicted using a reference set of alleles available at IEDB [54,55]. Low percentile epitopes are regarded as high affinity binders (Baldauf et al. 2015). Then, the binding affinity of the considered epitopes was predicted through a server MHCPred 2.0, and those fulfilling the criteria of the selection (IC50 values < 100 nM for DRB*0101) were chosen. DRB*0101 is an allele that is responsible for HLA susceptibility and is found in 95% of the population [55] The selected epitopes were further processed in order to check their virulence through VirulentPred [56]. The antigenicity of the virulent epitopes was again checked through VaxiJen 2.0. The AllerTop server was then used for the removal of the allergic epitopes [57].

2.4. Design of a Multi-Epitopes Vaccine

One of the main issues concerned with epitope vaccines is that they are weakly immunogenic, but this can be solved by joining multiple antigenic epitopes through linkers and designing a multi-epitopes vaccine [58]. Multi-epitopes vaccines are considered a better option to combat pathogenic infections [59]. The linkers used for fusing the epitopes were GPGPG [59]. The final construct was joined to the cholera toxin [60]. The tool Protparam [61] was then used for predicting the physicochemical properties of the vaccine. The 3D structure of the vaccine was predicted through 3D pro [62], and then loops of the vaccine were modeled using Galaxy loop [63], and refinement was conducted via Galaxy refinement [64]. Disulfide bonds were introduced via design 2.0 [65] to achieve stability of the final construct. Codon optimization and reverse translation of the vaccine candidate were achieved through the Jcat tool [66].

2.5. Simulating Host Immunity against Vaccine

Using C-immSim server, host immune responses were simulated against the designed vaccine candidate [67]. This server predicted the host immune system simulation in response to the vaccine in three different organs, namely the thymus, bone marrow, and the lymph nodes [68].

2.6. Docking and Refinement

To understand the binding affinity of the vaccine with the receptors of the immune system, a molecular docking study was performed [69]. This analysis is very crucial because the high affinity of the construct with the receptors means that the construct could generate good immune responses [70]. The vaccine construct was blindly docked to the MHC-I, MHC-II, and Toll-like Receptor 4 (TLR-4) [71]. TLR-4 helps in the production of cytokines, which leads to the activation of adaptive immunity. Molecular docking was carried out through a patchdock tool [72]. After docking, refinement of the docked complexes was carried out via Firedock [73]. After Firedock, only the complex with lowest energy was considered [73]. The intermolecular interactions of the complexes were interpreted through Chimera 1.13.1 [74].

2.7. Molecular Dynamic Analysis

The vaccine was evaluated in 300ns of a computer simulation. The purpose of this step was to decode the dynamics affinity of the vaccine construct for the immune receptors used. Furthermore, it helps to examine the epitopes presentation of the vaccine to the host immune system [75]. AMBER 20 was used for simulating the docked vaccine complexes [76]. In the first phase, an antechamber program was used to generate the complexes’ parameters [77]. The complexes were then solvated in a TIP3P solvation box (size 12 Å) [78] via the Leap program [79]. For the illustration of the intermolecular interactions of the systems, a ff14SB force field was applied [80]. Nine, eight, and seven Na+ counter ions were added for neutralization of the TLR4, MHC-I, and MHC-II complexes, respectively. The second phase was the preprocessing phase, which was about preparing the systems to be used in the production. The energy of the systems was optimized using the steepest descent and conjugate gradient algorithms. Systems were heated up to 300 K. Langavin dynamics [81] was used for a constant temperature, while a SHAKE algorithm was used to constrain the systems’ hydrogen bonds [82]. Moving further, the complexes were equilibrated for 100-ps. Pressure equilibration was conducted via an NPT ensemble [83]. Simulation trajectories of 300-ns were generated using the Berendsen algorithm [84]. A SHAKE algorithm was used to constrain hydrogen bonds, and trajectories analysis was performed through CPPTRAJ [85].

2.8. Calculation of TLR4-Vaccine Binding Energies

The Molecular Mechanics Poisson-Boltzmann Surface Area (MMPBSA) binding free energies for the vaccine-TLR4 were estimated through an AMBER20 MMPBSA.py module [86]. The Ante-MMPBSA.py module of AMBER was used for the prediction of the input parameter files of the complex, receptor, and vaccine. The binding energies of 100 different frames were determined, and the purpose was to determine the difference of free energy between the solvated and gas phases [87].

3. Results

A number of A. baumannii genome sequences are available in public databases, which can be retrieved for experimental and computational research works. In this study, we selected 50 complete sequenced A. baumannii genomes that consist of approximately 95,050 proteins.

3.1. Bacterial Pan-Genome Analysis

The pan-genome is the set of all genes present in a genome and contains a core genome (present in all strains), an accessory genome (those sequences that are present in two or more than two strains but not in all strains), and unique sequences (present only in a single strain) [88,89]. Bacterial pan-genome analysis was performed to obtain insights about the pathogen core genome. The A. baummanni strains have 95,050 core proteins (average number is 1901 proteins per genome). Accessory, unique, and exclusively absent proteins numbered 15,200, and 0, respectively [87]. The total number of proteins of each strain is graphically presented in Figure 2. The core-pan plot indicates that the strain’s pan-genome is in an open state, and chances are high of it gaining new genes in the future. Additionally, COG distribution found the core proteins involved in metabolic biogenesis [90].

3.2. CD HIT and PSORTB Analysis

The Cd-Hit is a server used for the identification of redundant and non-redundant sequences in a complete genome. Cd-hit analysis of the core sequences was performed to remove the redundant sequences. The redundant sequences were removed, and the non-redundant sequences were selected for further processing. CD hit analysis showed 1773 non-redundant proteins in the pathogen core proteome and were processed further. The number of redundant and non-redundant proteins is shown in Figure 3. Those proteins that are present on the surface of pathogen were considered as good vaccine candidates because they are exposed to the environment and have the potential to produce an immune response in the host. Therefore, proteins that are present in the extracellular, outer membrane, and periplasmic space are considered to be usable for vaccine designing. After PsortB analysis, 2 extracellular-, 11 outer membrane-, and 5 periplasmic proteins were unveiled. The localization of proteins is explained in Figure 3.

3.3. Antigenicity, Allergenicity, Human and Normal Microbiota Similarity, and Transmembrane Helices and Stability Analysis

All 18 filtered proteins were checked for antigenicity, allergenecity, and homology with human and normal flora genomes through Vaxijen, Allertop, and BLASTp, respectively. Antigenicity analysis predicted 13 proteins as antigenic with a score > 0.5. The Allertop 2.0 server found 4 protein sequences as allergic. Six proteins were homologous to the human genome, and 8 proteins were similar to three normal microbiota species. The non-homologous proteins reduce the chances of stimulating auto-immune reactions [91]. Likewise, no hits against probiotic bacteria demonstrated that beneficial microorganism growth will not be altered. Furthermore, transmembrane helices analysis revealed 0 proteins. Nine proteins were discarded as unstable (score > 40) because they had a molecular weight of >100 kDa (Figure 4). Out of 18 proteins, only 4 proteins were antigenic, non-allergen, and non-homologous. These proteins are: TonB-dependent siderophore receptor, two OmpA family proteins, and OprD family outer membrane porin; these were selected as the best vaccine candidates.

3.4. Physiochemical Properties

Different physicochemical properties of proteins can be predicted through protparam. The most important is molecular weight. Low molecular weight proteins (less than 110 kDa) have only such proteins as can be purified easily. The physiochemical properties of 18 virulence factor molecules are shown in Table 1.

3.5. Prediction of B Cell Epitopes

After performing the subtractive proteomic filters, 4 proteins were prioritized. B-cell and T-cell epitopes were predicted for all the selected vaccine candidates. The binding of antigen and antibody is necessary for making an antigen-antibody complex to stimulate the immune system. The adaptive immunity is specific in pathogen clearance [92]. Adaptive immunity converts B-cells into plasma cells that generate antibodies that recognize the pathogen on succeeding encounters [93]. These immunological responses are key in vaccination [94]. The four selected proteins—TonB-dependent siderophore receptor, two OmpA family protein, and OprD family outer membrane porin—were subjected to B-cell epitopes prediction. B-cell epitopes predicted for the mentioned proteins are in the following order: 15, 5, 4, and 11, respectively, as tabulated in Table S1. The B and T lymphocyte cells of the acquired immune cells are involved in provoking antibodies-dependent responses against invader microorganisms [95]. Thus, in this study, the final predicted B-cell epitopes were subjected to T-cell epitopes prediction. The selection of MHC-I and MHC-II epitopes [96] is based on the least percentile score as shown in Table S2. The human leukocyte antigen (HLA) system (MHC in humans) is a vital part of the human immune system. The HLA system is controlled by genes present on chromosome 6 and encode cell surface molecules that present antigenic peptides to the T-cell receptor (TCR) on T cells [97].

3.6. Epitope Filtration Phase

In the epitope filtration phase, predicted epitopes were subjected to evaluation of different checks, mainly including HLA DRB*0101 allele binding efficacy and antigenicity analysis. The HLA DRB*0101 gene is highly prevalent in the human population (95%) [98]. Epitopes of IC50 values < 100 nM for DRB*0101 were selected as they represent efficient immune responses stimulating epitopes (Table 2). The epitope filtration phase revealed all the predicted epitopes as probably antigenic, non-allergic, non-toxic, and water soluble [99], as shown in Table 2.

3.7. Multi-Epitopes Vaccine Construct Designing

The issue with epitope vaccines is that they are weakly immunogenic to induce an immune response, which can be resolved by joining the epitopes to make a multi-epitopes vaccine [100,101,102]. Adjutants and linkers are used to make a multi-epitopes vaccine. In total, 17 epitopes were selected based on their ability to clear all vaccine filters and were joined together by using GPGPG linkers. These linkers keep the epitopes separate and do not allow them to fold around each other. The designed vaccine construct is shown in Figure 5. The designed vaccine stability score and half-life is 33 and 10 h in E. coli, respectively.

3.8. 3D Structure Prediction

The construct was checked for stability, and the result showed that the designed vaccine construct is stable and can be used experimentally because of its small size. Subsequently, a 3D structure was predicted and is given in Figure 6.

3.9. Loops Modeling and Refinement

The stability of the vaccine construct is very important, and to increase the stability, loops modeling was performed to design an efficient 3D vaccine. Residues involved in loop formation, including, Met1-Val8, Ala19-Gly21, Cys30-Thr36, Glu50-Ile61, Thr62-Pro74, Gly185-Ile186, gly199-Gln200-Lys207, Arg202-Pro220, Gln221-Gly227, Gly228-Pro240, and Ser240-Phe220, were selected for loop modeling.

3.10. Disulfide Engineering

Disulfide bonds provide substantial stability to the protein. Disulfide engineering incorporates new disulfide bonds to replace highly unstable pair residues. Disulfide engineering was performed to further increase the stability of the folded vaccine construct by decreasing the conformational entropy. The intra and intra chains of the vaccine were checked for disulfide. Both original and mutant structures are shown in Figure 7. The following residue pairs were found to be highly unstable and were mutated. These residues are: Ala20Cys -Tyr25Cys, Cys27-Cys84, Thr36Cys-Cys95, His48Cys-Ser80Cys. Cys85-Cys110, Cys65-Gln69Cys, Cys65-Cys72, Ser90Cys -Ala96Cys, Lys104Cys -Ser123 Cys, Ile107Cys -Asp121Cys, and Val108Cys -Pro117Cys.

3.11. Codon Optimization

Codon optimization refers to a genetic approach to optimizing a given sequence as per the translation machinery of the host to obtain maximum expression of that exogenous sequence in the host expression system [103]. The codon optimization of the vaccine,, which is 0.92, was measured by a codon adaptation index (CAI); the GC content of the vaccine is 57.08%. These values indicate the efficient codon usage of the vaccine sequence in the E. coli K12 strain and hence it greater expression.

3.12. Molecular Docking

The interaction of the designed vaccine with the host receptors is necessary to stimulate the immune response against the designed vaccine. Blind docking of the vaccine was carried out with MHC-I, MHC-II, and TLR-4 receptors. The immune cell receptors were retrieved from an NCBI database using their specific PDB ID. The results of the blind docking are given in Tables S3–S5. In each case, 20 solutions were predicted, mainly the docking score.

3.13. Refinement of Docked Complexes

From the blind docking, the top 10 complexes were selected for refinement [59]. The top 10 solutions for each receptor are tabulated in Table 3, Table 4 and Table 5. Low global energy solutions were selected for binding conformation analysis [60]. The binding mode and interactions of the vaccine with TLR4, MHC-I, and MHC-II were studied. The selection of the top complex was carried out based on global binding energy. In the case of MHC-I, solution 7 was selected considering its global energy of −17.35 kcal/mol. For MHC-II, complex 8 was selected considering its global binding energy of −2.99 kcal/mol. For TLR-4, solution 2 was selected with a global binding energy of −3.22 kcal/mol. The docked intermolecular conformation of the vaccine to MHC-I, MHC-II, and TLR-4 is shown in Figure 8. It was also noticed that among the shortlisted solutions for each immune receptor were found different intermolecular conformations.

3.14. Chemical Interactions of the Vaccine with MHC-I, MHC-II, and TLR-4

Antigen presentation by MHC proteins is essential for acquired immunity. Prior to presentation, peptides must interact with various types of immune cells. These intermolecular interactions are critical to decipher, as they highlight the residues important from a vaccine recognition perspective. The designed vaccine showed robust interactions with several key hydrophilic and hydrophobic residues of receptor molecules, which are mentioned in Table 6.

3.15. Molecular Dynamic Simulation

The dynamics of the docked complexes were deciphered in order to make sure whether the binding of the vaccine to receptors was stable and whether the epitopes were in an exposed position to the host immune cells. The dynamics stability was evaluated through (i) root mean square deviation (RMSD), (ii) root mean square fluctuation (RMSF), and the (iii) radius of gyration (RoG) [104]. All these three types of analysis were conducted considering carbon alpha atoms. The RMSD plot of the systems was seen to steadily increase with no major deviations seen. This implicates the stable intermolecular binding between vaccine and receptors throughout the length of the simulation time. This also confers the continuous presentation of vaccine epitopes to the immune system for activation of humoral and cellular immunity. The maximum RMSD for each system reaches 10 Å (Figure 9A). The higher RMSD is the result of a large vaccine-receptors complex, but despite that, the vaccine is strongly held at the docked site of the receptors. Second, the RMSF was calculated to determine the residue level fluctuations of receptor molecules (Figure 9B). This was vital, as the presence of vaccine either heightens the residues; stability or makes it highly unstable. Most of the receptor residues show good stability (<3 Å). The loops are flexible, and in the presence of the vaccine, such loops become highly flexible, showing a high RMSF. Lastly, RoG assay was conducted to examine the complexes’ compactness with respect to time (Figure 9C). The RoG complemented the RMSD and concluded that the systems are in an equilibrium state.

3.16. Calculation of Vaccine-Receptors Binding Energies

Binding free energies of the docked molecules were estimated through MM-GBSA, and MM-PBSA techniques were used for validation of the binding efficacy of the docked complexes. The total binding free energies of the MHC-I, MHC-II, and TLR-4-vaccine complex were −178.62 kcal/mol, −181.94 kcal/mol, and −169.83 kcal/mol, respectively (Table 7). A major contributor to the binding net energy came from the electrostatic energy as well as van der Waals energy, and a non-favorable (not contributing to the overall stability of the complexes) contribution was seen from solvation energy.

3.17. Immune Simulation of the Designed Vaccine

In this analysis, the host response against the modeled vaccine was predicted for 350 days and it was analyzed that there is a high production of IgG and IgM antibodies against the designed vaccine. The secondary response was followed by a tertiary immune response, which led to the high-level production of B-cells and “IgM + IgG, IgM, IgG1 + IgG2, IgG1 and IgG2” (Figure 10A). An increase in interferon was also detected for 350 days, and the results are mentioned in Figure 10B. The different B-cell and T-cell immune responses in response to the antigen are given in Figures S1 and S2.

4. Discussion

We are moving towards the end of the antibiotic era. The crisis of AR and new bacterial strains are emerging. Some bacterial species have already been declared as resistant to all available antibiotics [105]. A. baumannii is one of those species of bacteria that has been reported as resistant to all available antibiotics [106]. A. baumannii is also capable of adopting new mechanisms of resistance and is evolving very quickly [107]. A. baumannii, is also able to develop a quick resistance to antibiotics due to its plastic genome and has the capability of causing healthcare associated infections and has the potential to survive in highly detrimental conditions [108]. The growing burden of microbial resistance has a remarkable clinical and economical effect and demands the development of new therapeutic candidates to treat infections of bacterial pathogens.
Vaccines have the potential to provide a permanent solution to tackling AR bacteria. RV, which is a genome-based vaccine development technique, is a popular approach in identifying new vaccine candidates [109]. RV was developed by Dr. Rino Rappuoli and is an emerging in silico vaccine development framework [109,110]. In the recent past, RV has contributed significantly to vaccine development against pathogens that are unable to be addressed by Pasteur’s principles of vaccinology [17]. The traditional Pasteur vaccinology approach is not an appropriate choice for pathogens that cannot be cultured [111]. Likewise, it ignores conservation and molecular mimicry of antigenic determinants [79], as surfaced in vaccine development against Neisseria and Mycobacterium species [112]. The choice of subunit vaccines is also very costly and time consuming, and the likelihood of screening potential antigenic epitopes is low [113]. The availability of an exponential amount of genomic data in public databases and advancement in bioinformatics tools have significantly sped up the vaccine development process [114]. Vaccine candidates can be identified at the pathogen surface through RV [112], which has been successfully used in meningococci serogroup B vaccine (4CMenB) development [111]. RV has been extensively exploited for designing vaccines against many bacterial and viral pathogens [35]. Lai et al. used glycoprotein from Ebolavirus and designed an MHC-I epitope-based vaccine by employing an RV approach and subsequently obtained promising results for a proposed vaccine via experimental validation (https://doi.org/10.3389/fmicb.2017.01571 accessed on 22 April 2022). In another study, Agallou et al. performed mice experiments of eight MHC-I and MHC-II epitopes predicted using an RV approach and proposed that these epitopes can provide better immunity against Leishmania infantum (https://doi.org/10.3389/fimmu.2014.00268 accessed on 22 April 2022). Additionally, Singh et al. predicted FilF as a potential vaccine target using RV and immunoinformatics approaches and validated its 50% immune protective efficacy in a murine pneumonia model (https://doi:10.3389/fmicb.2016.00158 accessed on 22 April 2022). In recent advancements, classical reverse vaccinology is integrated with pan-genomic-based reverse vaccinology (PGRV) to identify core genome antigens. PGRV is successfully used to map four protective antigens in Streptococcus agalactiae genomes [114,115]. There is no vaccine in the development process for A. baumannii, which makes the situation more worrisome. Therefore, substantial efforts were needed to identify protective antigens from the core genome of the pathogen, which can easily be analyzed in experimental studies.
The current study was about computational prioritization of vaccine candidates against A. baumannii based on RV and pan-genome analysis. In the current study, we report four outer membrane proteins: TonB-dependent siderphore receptor, OmpA family protein, type IV pilus biogenesis stability protein, and OprD family outer membrane porin, all of which highly susceptible to an effective vaccine. The OmpA is a well conserved protein among Acinetobacter species and can produce broad-range immunity. Besides this, OmpA has a soluble and stable structure, making it an attractive vaccine target. This protein also has favorable adhesion and haemagglutinins properties. However, as the OmpA pure form is insoluble, its delivery is hard (https://doi.org/10.22038/ijbms.2019.30799.7427 accessed on 22 April 2022). The combination of OmpA antigenic epitopes with those of the TonB-dependent siderphore receptor, type IV pilus biogenesis stability protein, and OprD family outer membrane porin might generate an ideal vaccine construct capable of generating strong and protective immune responses. The findings of the study will save time and may lead to cost effective vaccine development. We used RV in combination with immunoinformatics and biophysical approaches to construct a chimeric vaccine against superbug A. baumannii and understand its binding potential with the host MHC-I, MHC-II, and TLR-4 receptors to examine its immune system presentation and ability to confer protection against the antigen. At the end, a host immune system simulation was performed against the vaccine in order to predict the antibodies’ production and cellular immune responses. This in silico simulation has many limitations, for instance, it does not include type I IFN, which has been recently evident as potentially important in the management of infection and can be further improved using more experimental knowledge. To date, many attempts have been made to develop vaccines against A. baumannii, yet there is no successful vaccine against this pathogen that has been identified [116,117,118]. Several attempts were also made to form a live attenuated vaccine against A. baumannii, and recently a D-glutamine strain of pathogen was used to make a live attenuated vaccine. A strain of mutated morl1 and morl2 genes was synthesized. Although it induced both antibody and cell mediated immunity, the genes were less virulent and the response was weak [119].

5. Concluding Remarks and Limitations

In this in silico study, several different bioinformatics techniques, ranging from subtractive proteomics to immunoinformatics techniques, were considered, with the aim of designing a multi-epitopes-based vaccine against a nosocomial superbug A. baumannii bacteria. These techniques have been applied to design vaccines against different pathogens [116,117,118]. The rationale behind this work is the lack of an FDA approved vaccine for A. baumannii, and the pathogen shows a high level of resistance to many available antibiotics [120]. The designed vaccine construct comprises only antigenic epitopes that do not lack allergic sequences, but it is capable of eliciting strong immune responses. The epitopes were filtered from four vaccine proteins prioritized using vaccine candidacy criteria. These proteins are: TonB-dependent siderphore receptor, OmpA family protein, type IV pilus biogenesis stability protein, and OprD family outer membrane porin. The designed vaccine construct has shown excellent binding potency to the immune cell’s receptors, disclosed proper binding confirmation, and produced strong binding energies. We believe that the findings and predictions of our study may accelerate the vaccine development process against A. baumannii and may deliver data to speed up vaccine development against the pathogen. Moreover, the findings of the study will also save time and cost and will be helpful for vaccinologists in developing a vaccine against A. baumannii. Although we were very careful in the selection process of epitopes, the study has some limitations. The vaccine construct epitopes’ proper fusion needs thorough experimental testing, which is not conducted herein. The predictions made in the study are based on tools and servers. These predictions are not that accurate due to the lack of healthy experimental data to train the tools and servers used. Lastly, the real immune protection of the designed vaccine required a wide range of in vivo and in vitro testing studies.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijerph19095568/s1. Table S1. List of B cell epitopes predicted from four potential vaccine proteins. Table S2. MHC –I and MHC-II predicted epitopes. Table S3. Docking score of top 20 complexes of designed vaccine construct to MHC-I complexes generated by patchDock server. Table S4. Docking score of top 20 complexes of designed vaccine construct to MHC-II complexes generated by patchDock server. Table S5. Docking score of top 20 complexes of designed vaccine construct to TLR4 complexes generated by patchDock server. Figure S1. B cell and T cell population produced in response to the vaccine antigen. Figure S2. Immune cell response generated in response to chimeric vaccine construct. Tc (cytotoxic killer T-cell) Macrophages (Mφ) Nature killer cell, Dendritic and epithelial cell.

Author Contributions

Conceptualization, A.A., S.A. and M.T.u.Q.; Data curation, M.u.-d., A.A., A.U., S.I., A.N., M.K., M.u.H., Z.A., Y.E.B., M.S. and B.S.; Formal analysis, M.u.-d., A.A. and S.A.; Funding acquisition, A.A.; Investigation, M.u.-d., A.A., S.A. and M.T.u.Q.; Methodology, A.U., S.I., A.N., M.K., M.u.H., Z.A., Y.E.B., M.S. and B.S.; Project administration, A.A., S.A. and M.T.u.Q.; Resources, A.A.; Software, S.A. and M.T.u.Q.; Supervision, S.A.; Validation, A.U., S.I., S.A., A.N., M.K., M.u.H., Z.A., Y.E.B., M.S., B.S. and M.T.u.Q.; Visualization, M.u.-d. and A.A.; Writing—original draft, M.u.-d. and A.A.; Writing—review & editing, A.U., S.I., S.A., A.N., M.K., M.u.H., Z.A., Y.E.B., M.S., B.S. and M.T.u.Q. 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

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. Alanis, A.J. Resistance to antibiotics: Are we in the post-antibiotic era? Arch. Med. Res. 2005, 36, 697–705. [Google Scholar] [CrossRef] [PubMed]
  2. Chokshi, A.; Sifri, Z.; Cennimo, D.; Horng, H. Global contributors to antibiotic resistance. J. Glob. Infect. Dis. 2019, 11, 36. [Google Scholar] [PubMed]
  3. Kulik, E.M.; Thurnheer, T.; Karygianni, L.; Walter, C.; Sculean, A.; Eick, S. Antibiotic susceptibility patterns of aggregatibacter actinomycetemcomitans and porphyromonas gingivalis strains from different decades. Antibiotics 2019, 8, 253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. PCAST. National Action Plan for Combatting Antibiotic-Resistant Bacteria; White House: Washington, DC, USA, 2015.
  5. Ventola, C.L. The antibiotic resistance crisis: Part 2: Management strategies and new agents. Pharm. Ther. 2015, 40, 344. [Google Scholar]
  6. Brooks, B.D.; Brooks, A.E. Therapeutic strategies to combat antibiotic resistance. Adv. Drug Deliv. Rev. 2014, 78, 14–27. [Google Scholar] [CrossRef] [PubMed]
  7. Annunziato, G. Strategies to Overcome Antimicrobial Resistance (AMR) Making Use of Non-Essential Target Inhibitors: A Review. Int. J. Mol. Sci. 2019, 20, 5844. [Google Scholar] [CrossRef] [Green Version]
  8. Yanfen, D.; Li, T.; Wan, Y.; Liao, P. Signal Molecule-Dependent Quorum-Sensing and Quorum-Quenching Enzymes in Bacteria. Crit. Rev. Eukaryot. Gene Expr. 2014, 24, 117–132. [Google Scholar]
  9. Hall, N. Advanced Sequencing Technologies and Their Wider Impact in Microbiology. J. Exp. Biol. 2007, 210, 1518–1525. [Google Scholar] [CrossRef] [Green Version]
  10. Afreenish, H.; Usman, J.; Kaleem, F.; Khan, A.; Hussain, Z. In Vitro Activity of Aminoglycosides, Lactam-Lactamases Inhibitor Combinations and Tetracyclines against Multi-Drug Resistant Acinetobacter Baumannii, Isolated from a Tertiary Care Hospital. J. Microbiol. Antimicrob. 2010, 2, 47–50. [Google Scholar]
  11. Kadri, S.S. Key Takeaways from the US CDC’s 2019 Antibiotic Resistance Threats Report for Frontline Providers. Crit. Care Med. 2020, 48, 939–945. [Google Scholar]
  12. Emmett, K.; Perkins, M.D.; Small, P.; Hanson, C.; Reed, S.; Cunningham, J.; Aledort, J.E.; Hillborne, L.; Rafael, M.E.; Girosi, F. Reducing the Global Burden of Tuberculosis: The Contribution of Improved Diagnostics. Nature 2006, 444, 49–57. [Google Scholar]
  13. Mobarki, N.; Almerabi, B.; Hattan, A. Antibiotic Resistance Crisis. Int. J. Med. Dev. Ctries 2019, 40, 561–564. [Google Scholar] [CrossRef] [Green Version]
  14. Oved, K. Fighting AMR with Host Immune Response Technology. Drug Discov. Today 2021, 26, 2081. [Google Scholar] [CrossRef] [PubMed]
  15. ProtParam. Available online: https://bio.tools/protparam (accessed on 21 April 2022).
  16. Mousumi, S.; Sarkar, A. Review on Multiple Facets of Drug Resistance: A Rising Challenge in the 21st Century. J. Xenobiotics 2021, 11, 197–214. [Google Scholar]
  17. Taylor, D. The Pharmaceutical Industry and the Future of Drug Development. 2015. Available online: https://pubs.rsc.org/en/content/chapterhtml/2015/bk9781782621898-00001?isbn=978-1-78262-189-8 (accessed on 22 April 2022).
  18. Lantian, Z.; Zhang, Y.; Ge, Y.; Zhu, X.; Pan, J. Regulatory Mechanisms and Promising Applications of Quorum Sensing-Inhibiting Agents in Control of Bacterial Biofilm Formation. Front. Microbiol. 2020, 2558. [Google Scholar] [CrossRef]
  19. Aslam, B.; Wang, W.; Arshad, M.I.; Khurshid, M.; Muzammil, S.; Rasool, M.H.; Nisar, M.A.; Alvi, R.F.; Aslam, M.A.; Qamar, M.U. Antibiotic resistance: A rundown of a global crisis. Infect. Drug Resist. 2018, 11, 1645. [Google Scholar] [CrossRef] [Green Version]
  20. Kennedy, D.A.; Read, A.F.; Read, A.F. Why does drug resistance readily evolve but vaccine resistance does not? Proc. R. Soc. B Biol. Sci. 2017, 284, 20162562. [Google Scholar] [CrossRef]
  21. Goldsby, R.A.; Kindt, T.J.; Osborne, B.A.; Kuby, J. Chapter 2: Cells and Organs of the Immune System. In Immunology, 5th ed.; W. H. Freeman and Company: New York, NY, USA, 2003; pp. 24–56. Available online: https://www.scirp.org/(S(351jmbntvnsjt1aadkposzje))/reference/ReferencesPapers.aspx?ReferenceID=1913594 (accessed on 22 April 2022).
  22. Plotkin, S. History of vaccination. Proc. Natl. Acad. Sci. USA 2014, 111, 12283–12287. [Google Scholar] [CrossRef] [Green Version]
  23. Moriel, D.G.; Scarselli, M.; Serino, L.; Mora, M.; Rappuoli, R.; Masignani, V. Genome-based vaccine development: A short cut for the future. Hum. Vaccines 2008, 4, 184–188. [Google Scholar] [CrossRef] [Green Version]
  24. Baseer, S.; Ahmad, S.; Ranaghan, K.E.; Azam, S.S. Towards a peptide-based vaccine against Shigella sonnei: A subtractive reverse vaccinology based approach. Biologicals 2017, 50, 87–99. [Google Scholar] [CrossRef] [Green Version]
  25. Eliopoulos, G.M.; Maragakis, L.L.; Perl, T.M. Acinetobacter baumannii: Epidemiology, antimicrobial resistance, and treatment options. Clin. Infect. Dis. 2008, 46, 1254–1263. [Google Scholar]
  26. Gellings, P.S.; Wilkins, A.A.; Morici, L.A. Recent Advances in the Pursuit of an Effective Acinetobacter baumannii Vaccine. Pathogens 2020, 9, 1066. [Google Scholar] [CrossRef] [PubMed]
  27. 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]
  28. 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]
  29. 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] [PubMed]
  30. Kumar Jaiswal, A.; Tiwari, S.; Jamal, S.B.; Barh, D.; Azevedo, V.; Soares, S.C. An in silico identification of common putative vaccine candidates against treponema pallidum: A reverse vaccinology and subtractive genomics based approach. Int. J. Mol. Sci. 2017, 18, 402. [Google Scholar] [CrossRef]
  31. Johri, S.; Solanki, J.; Cantu, V.A.; Fellows, S.R.; Edwards, R.A.; Moreno, I.; Vyas, A.; Dinsdale, E.A. ‘Genome skimming’ with the MinION hand-held sequencer identifies CITES-listed shark species in India’s exports market. Sci. Rep. 2019, 9, 4476. [Google Scholar] [CrossRef]
  32. Butt, A.M.; Tahir, S.; Nasrullah, I.; Idrees, M.; Lu, J.; Tong, Y. Mycoplasma genitalium: A comparative genomics study of metabolic pathways for the identification of drug and vaccine targets. Infect. Genet. Evol. 2012, 12, 53–62. [Google Scholar] [CrossRef]
  33. Sikic, K.; Carugo, O. Protein sequence redundancy reduction: Comparison of various method. Bioinformation 2010, 5, 234. [Google Scholar] [CrossRef] [Green Version]
  34. 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]
  35. Rizwan, M.; Naz, A.; Ahmad, J.; Naz, K.; Obaid, A.; Parveen, T.; Ahsan, M.; Ali, A. VacSol: A high throughput in silico pipeline to predict potential therapeutic targets in prokaryotic pathogens using subtractive reverse vaccinology. BMC Bioinform. 2017, 18, 106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. 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] [PubMed]
  37. Ahmad, S.; Ranaghan, K.E.; Azam, S.S. Combating tigecycline resistant Acinetobacter baumannii: A leap forward towards multi-epitope based vaccine discovery. Eur. J. Pharm. Sci. 2019, 132, 1–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Hara, Y.; Mohamed, R.; Nathan, S. Immunogenic Burkholderia pseudomallei outer membrane proteins as potential candidate vaccine targets. PLoS ONE 2009, 4, e6496. [Google Scholar] [CrossRef]
  39. Collins, B.S. Gram-negative outer membrane vesicles in vaccine development. Discov. Med. 2011, 12, 7–15. [Google Scholar]
  40. Rashid, M.I.; Naz, A.; Ali, A.; Andleeb, S. Prediction of vaccine candidates against Pseudomonas aeruginosa: An integrated genomics and proteomics approach. Genomics 2017, 109, 274–283. [Google Scholar] [CrossRef]
  41. Chen, L.; Yang, J.; Yu, J.; Yao, Z.; Sun, L.; Shen, Y.; Jin, Q. VFDB: A reference database for bacterial virulence factors. Nucleic Acids Res. 2005, 33, D325–D328. [Google Scholar] [CrossRef] [Green Version]
  42. Kaushik, D.K.; Sehgal, D. Developing antibacterial vaccines in genomics and proteomics era. Scand. J. Immunol. 2008, 67, 544–552. [Google Scholar] [CrossRef]
  43. 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]
  44. ExPASy. Available online: https://www.expasy.org/ (accessed on 22 April 2022).
  45. Guruprasad, K.; Reddy, B.V.B.; Pandit, M.W. Correlation between stability of a protein and its dipeptide composition: A novel approach for predicting in vivo stability of a protein from its primary sequence. Protein Eng. Des. Sel. 1990, 4, 155–161. [Google Scholar] [CrossRef]
  46. Tusnady, G.E.; Simon, I. The HMMTOP transmembrane topology prediction server. Bioinformatics 2001, 17, 849–850. [Google Scholar] [CrossRef] [PubMed]
  47. 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] [PubMed] [Green Version]
  48. Sachdeva, G.; Kumar, K.; Jain, P.; Ramachandran, S.; Ramachandran, S. SPAAN: A software for prediction of adhesins and adhesin-like proteins using neural networks. Bioinformatics 2004, 16, 16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Wadood, A.; Jamal, A.; Riaz, M.; Khan, A.; Uddin, R.; Jelani, M.; Azam, S.S. Subtractive genome analysis for in silico identification and characterization of novel drug targets in Streptococcus pneumonia strain JJA. Microb. Pathog. 2018, 115, 194–198. [Google Scholar] [CrossRef]
  50. He, Y.; Xiang, Z.; Mobley, H.L.T. Vaxign: The first web-based vaccine design program for reverse vaccinology and applications for vaccine development. J. Biomed. Biotechnol. 2010, 2010, 297505. [Google Scholar] [CrossRef]
  51. 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]
  52. Naz, K.; Naz, A.; Ashraf, S.T.; Rizwan, M.; Ahmad, J.; Baumbach, J.; Ali, A. PanRV: Pangenome-reverse vaccinology approach for identifications of potential vaccine candidates in microbial pangenome. BMC Bioinform. 2019, 20, 123. [Google Scholar] [CrossRef]
  53. Hassan, A.; Naz, A.; Obaid, A.; Paracha, R.Z.; Naz, K.; Awan, F.M.; Muhmmad, S.A.; Janjua, H.A.; Ahmad, J.; Ali, A. Pangenome and immuno-proteomics analysis of Acinetobacter baumannii strains revealed the core peptide vaccine targets. BMC Genom. 2016, 17, 732. [Google Scholar] [CrossRef] [Green Version]
  54. Vita, R.; Overton, J.A.; Greenbaum, J.A.; Ponomarenko, J.; Clark, J.D.; Cantrell, J.R.; Wheeler, D.K.; Gabbard, J.L.; Hix, D.; Sette, A. The immune epitope database (IEDB) 3.0. Nucleic Acids Res. 2015, 43, D405–D412. [Google Scholar] [CrossRef]
  55. Janeway, C.A., Jr.; Travers, P.; Walport, M.; Shlomchik, M.J. The major histocompatibility complex and its functions. In Immunobiology: The Immune System in Health and Disease, 5th ed.; Garland Science: New York, NY, USA, 2001. [Google Scholar]
  56. Garg, A.; Gupta, D. VirulentPred: A SVM based prediction method for virulent proteins in bacterial pathogens. BMC Bioinform. 2008, 9, 62. [Google Scholar] [CrossRef] [Green Version]
  57. Dimitrov, I.; Flower, D.R.; Doytchinova, I. AllerTOP-a server for in silico prediction of allergens. BMC Bioinform. 2013, 14, S4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. 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]
  59. Skwarczynski, M.; Toth, I. Peptide-based synthetic vaccines. Chem. Sci. 2016, 7, 842–854. [Google Scholar] [CrossRef] [Green Version]
  60. Baldauf, K.J.; Royal, J.; Hamorsky, K.T.; Matoba, N. Cholera Toxin B: One Subunit with Many Pharmaceutical Applications. Toxins 2015, 7, 974–996. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. ProtParam, K.E. ExPASy-ProtParam Tool. 2017. Available online: https://web.expasy.org/protparam/ (accessed on 1 April 2022).
  62. Cheng, J.; Randall, A.Z.; Sweredoski, M.J.; Baldi, P. SCRATCH: A protein structure and structural feature prediction server. Nucleic Acids Res. 2005, 33, W72–W76. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Giardine, B.; Riemer, C.; Hardison, R.C.; Burhans, R.; Elnitski, L.; Shah, P.; Zhang, Y.; Blankenberg, D.; Albert, I.; Taylor, J.; et al. Galaxy: A platform for interactive large-scale genome analysis. Genome Res. 2005, 15, 1451–1455. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Heo, L.; Park, H.; Seok, C. GalaxyRefine: Protein structure refinement driven by side-chain repacking. Nucleic Acids Res. 2013, 41, W384–W388. [Google Scholar] [CrossRef] [Green Version]
  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. Grote, A.; Hiller, K.; Scheer, M.; Münch, R.; Nörtemann, B.; Hempel, D.C.; Jahn, D. JCat: A novel tool to adapt codon usage of a target gene to its potential expression host. Nucleic Acids Res. 2005, 33, W526–W531. [Google Scholar] [CrossRef]
  67. Rapin, N.; Lund, O.; Castiglione, F. C-Immsim 10.1 Server. 2012. Available online: https://www.iac.cnr.it/~filippo/projects/c-immsim-online.html (accessed on 22 April 2022).
  68. Rapin, N.; Lund, O.; Bernaschi, M.; Castiglione, F. Computational immunology meets bioinformatics: The use of prediction tools for molecular binding in the simulation of the immune system. PLoS ONE 2010, 5, e9862. [Google Scholar] [CrossRef] [Green Version]
  69. Morris, G.M.; Lim-Wilby, M. Molecular docking. In Molecular Modeling of Proteins; Springer: Berlin/Heidelberg, Germany, 2008; pp. 365–382. [Google Scholar]
  70. 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] [PubMed]
  71. Ohto, U.; Yamakawa, N.; Akashi-Takamura, S.; Miyake, K.; Shimizu, T. Structural analyses of human Toll-like receptor 4 polymorphisms D299G and T399I. J. Biol. Chem. 2012, 287, 40611–40617. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Schneidman-Duhovny, D.; Inbar, Y.; Nussinov, R.; Wolfson, H.J. PatchDock and SymmDock: Servers for rigid and symmetric docking. Nucleic Acids Res. 2005, 33, W363–W367. [Google Scholar] [CrossRef] [Green Version]
  73. Andrusier, N.; Nussinov, R.; Wolfson, H.J. FireDock: Fast interaction refinement in molecular docking. Proteins Struct. Funct. Bioinforma. 2007, 69, 139–159. [Google Scholar] [CrossRef] [PubMed]
  74. 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] [PubMed] [Green Version]
  75. Andleeb, S.; Imtiaz-Ud-Din; Rauf, M.K.; Azam, S.S.; Badshah, A.; Sadaf, H.; Raheel, A.; Tahir, M.N.; Raza, S. A one-pot multicomponent facile synthesis of dihydropyrimidin-2(1: H)-thione derivatives using triphenylgermane as a catalyst and its binding pattern validation. RSC Adv. 2016, 6, 79651–79661. [Google Scholar] [CrossRef]
  76. Case, D.A.; Cerutti, D.S.; Cheateham, T.E.; Darden, T.A.; Duke, R.E.; Giese, T.J.; Gohlke, H.; Goetz, A.W.; Greene, D.; Homeyer, N.; et al. AMBER16 Package, University of California, San Francisco. 2016. Available online: https://ambermd.org/doc12/Amber16.pdf (accessed on 22 April 2022).
  77. Wang, J.; Wang, W.; Kollman, P.A.; Case, D.A. Antechamber: An accessory software package for molecular mechanical calculations. J. Am. Chem. Soc. 2001, 222, U403. [Google Scholar]
  78. Brice, A.R.; Dominy, B.N. Examining electrostatic influences on base-flipping: A comparison of TIP3P and GB solvent models. Commun. Comput. Phys. 2013, 13, 223–237. [Google Scholar] [CrossRef]
  79. Kerrigan, J.E. AMBER 10.0 Introductory Tutorial. 2009. Available online: https://ambermd.org/ (accessed on 1 April 2022).
  80. Case, D.A.; Babin, V.; Berryman, J.T.; Betz, R.M.; Cai, Q.; Cerutti, D.S.; Cheatham, T.E., III; Darden, T.A.; Duke, R.E.; Gohlke, H.; et al. The FF14SB force field. Amber 2014, 14, 29–31. [Google Scholar]
  81. Lavenda, B.H. Statistical Physics: A Probabilistic Approach; Courier Dover Publications: Mineola, NY, USA, 2016; ISBN 0486810313. [Google Scholar]
  82. Case, D.A.; Cheatham, T.E., III; Darden, T.; Gohlke, H.; Luo, R.; Merz, K.M., Jr.; Onufriev, A.; Simmerling, C.; Wang, B.; Woods, R.J. The Amber biomolecular simulation programs. J. Comput. Chem. 2005, 26, 1668–1688. [Google Scholar] [CrossRef] [Green Version]
  83. Feller, S.E.; Zhang, Y.; Pastor, R.W.; Brooks, B.R. Constant pressure molecular dynamics simulation: The Langevin piston method. J. Chem. Phys. 1995, 103, 4613–4621. [Google Scholar] [CrossRef]
  84. Lemak, A.S.; Balabaev, N.K. On the Berendsen thermostat. Mol. Simul. 1994, 13, 177–187. [Google Scholar] [CrossRef]
  85. 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]
  86. 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]
  87. Genheden, S.; Ryde, U. The MM/PBSA and MM/GBSA methods to estimate ligand-binding affinities. Expert Opin. Drug Discov. 2015, 10, 449–461. [Google Scholar] [CrossRef]
  88. Tahir ul Qamar, M.; Zhu, X.; Khan, M.S.; Xing, F.; Chen, L.L. Pan-genome: A promising resource for noncoding RNA discovery in plants. Plant Genome 2020, 13, e20046. [Google Scholar] [CrossRef]
  89. Ul Qamar, M.T.; Zhu, X.; Xing, F.; Chen, L.-L.L.; Tahir Ul Qamar, M.; Zhu, X.; Xing, F.; Chen, L.-L.L. ppsPCP: A plant presence/absence variants scanner and pan-genome construction pipeline. Bioinformatics 2019, 35, 4156–4158. [Google Scholar] [CrossRef]
  90. Araújo, C.L.; Alves, J.; Nogueira, W.; Pereira, L.C.; Gomide, A.C.; Ramos, R.; Azevedo, V.; Silva, A.; Folador, A. Prediction of new vaccine targets in the core genome of Corynebacterium pseudotuberculosis through omics approaches and reverse vaccinology. Gene 2019, 702, 36–45. [Google Scholar] [CrossRef]
  91. Galperin, M.Y.; Kristensen, D.M.; Makarova, K.S.; Wolf, Y.I.; Koonin, E. V Microbial genome analysis: The COG approach. Brief. Bioinform. 2019, 20, 1063–1070. [Google Scholar] [CrossRef]
  92. Kardani, K.; Bolhassani, A.; Namvar, A. An overview of in silico vaccine design against different pathogens and cancer. Expert Rev. Vaccines 2020, 19, 699–726. [Google Scholar] [CrossRef]
  93. Dorosti, H.; Eslami, M.; Negahdaripour, M.; Ghoshoon, M.B.; Gholami, A.; Heidari, R.; Dehshahri, A.; Erfani, N.; Nezafat, N.; Ghasemi, Y. Vaccinomics approach for developing multi-epitope peptide pneumococcal vaccine. J. Biomol. Struct. Dyn. 2019, 37, 3524–3535. [Google Scholar] [CrossRef] [PubMed]
  94. Bragazzi, N.L.; Gianfredi, V.; Villarini, M.; Rosselli, R.; Nasr, A.; Hussein, A.; Martini, M.; Behzadifar, M. Vaccines meet big data: State-ofthe- Art and future prospects. From the classical 3is (“isolate-inactivate-inject”) vaccinology 1.0 to vaccinology 3.0, vaccinomics, and Beyond: A historical overview. Front. Public Health 2018, 6, 62. [Google Scholar] [CrossRef] [PubMed]
  95. Walker, B.D.; Burton, D.R. Toward an AIDS vaccine. Science 2008, 320, 760–764. [Google Scholar] [CrossRef] [PubMed]
  96. Longbrake, E.E.; Cantoni, C.; Chahin, S.; Cignarella, F.; Cross, A.H.; Piccio, L. Dimethyl fumarate induces changes in B-and T-lymphocyte function independent of the effects on absolute lymphocyte count. Mult. Scler. J. 2018, 24, 728–738. [Google Scholar] [CrossRef]
  97. Ahmad, B.; Ashfaq, U.A.; Rahman, M.; Masoud, M.S.; Yousaf, M.Z. Conserved B and T cell epitopes prediction of ebola virus glycoprotein for vaccine development: An immuno-informatics approach. Microb. Pathog. 2019, 132, 243–253. [Google Scholar] [CrossRef]
  98. Choga, W.-T.; Anderson, M.; Phinius, B.-B.; Mbangiwa, T.; Bell, T.-G.; Seatla, K.-K.; Musonda, R.-M.; Moyo, S.; Blackard, J.-T.; Gaseitsiwe, S. A25 Impact of polymorphism in the hepatitis B surface gene on human leukocyte antigen (HLA) class II. Virus Evol. 2019, 5, vez002–vez024. [Google Scholar] [CrossRef]
  99. 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]
  100. Ul Qamar, M.T.; Bari, A.; Adeel, M.M.; Maryam, A.; Ashfaq, U.A.; Du, X.; Muneer, I.; Ahmad, H.I.; Wang, J.; Tahir Ul Qamar, M.; et al. Peptide vaccine against chikungunya virus: Immuno-informatics combined with molecular docking approach. J. Transl. Med. 2018, 16, 298. [Google Scholar] [CrossRef]
  101. 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]
  102. 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]
  103. 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] [PubMed]
  104. Jaydari, A.; Nazifi, N.; Forouharmehr, A. Computational design of a novel multi-epitope vaccine against Coxiella burnetii. Hum. Immunol. 2020, 81, 596–605. [Google Scholar] [CrossRef] [PubMed]
  105. MacLean, R.C.; San Millan, A. The evolution of antibiotic resistance. Science 2019, 365, 1082–1083. [Google Scholar] [CrossRef] [PubMed]
  106. Peleg, A.Y.; Seifert, H.; Paterson, D.L. Acinetobacter baumannii: Emergence of a successful pathogen. Clin. Microbiol. Rev. 2008, 21, 538–582. [Google Scholar] [CrossRef] [Green Version]
  107. Dijkshoorn, L.; Nemec, A.; Seifert, H. An increasing threat in hospitals: Multidrug-Resistant Acinetobacter baumannii. Nat. Rev. Microbiol. 2007, 5, 939–951. [Google Scholar] [CrossRef]
  108. Wareth, G.; Brandt, C.; Sprague, L.D.; Neubauer, H.; Pletz, M.W. Spatio-temporal distribution of Acinetobacter baumannii in Germany—A comprehensive systematic review of studies on resistance development in humans (2000–2018). Microorganisms 2020, 8, 375. [Google Scholar] [CrossRef] [Green Version]
  109. Bambini, S.; Rappuoli, R. The use of genomics in microbial vaccine development. Drug Discov. Today 2009, 14, 252–260. [Google Scholar] [CrossRef]
  110. 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. [Google Scholar] [CrossRef] [Green Version]
  111. Mora, M.; Veggi, D.; Santini, L.; Pizza, M.; Rappuoli, R. Reverse vaccinology. Drug Discov. Today 2003, 8, 459–464. [Google Scholar] [CrossRef]
  112. Adu-Bobie, J.; Capecchi, B.; Serruto, D.; Rappuoli, R.; Pizza, M. Two years into reverse vaccinology. Vaccine 2003, 21, 605–610. [Google Scholar] [CrossRef]
  113. Vaishnav, N.; Gupta, A.; Paul, S.; John, G.J. Overview of computational vaccinology: Vaccine development through information technology. J. Appl. Genet. 2015, 56, 381–391. [Google Scholar] [CrossRef] [PubMed]
  114. Albutti, A. An integrated computational framework to design a multi-epitopes vaccine against Mycobacterium tuberculosis. Sci. Rep. 2021, 11, 21929. [Google Scholar] [CrossRef] [PubMed]
  115. Maione, D.; Margarit, I.; Rinaudo, C.D.; Masignani, V.; Mora, M.; Scarselli, M.; Tettelin, H.; Brettoni, C.; Iacobini, E.T.; Rosini, R. Identification of a universal Group B streptococcus vaccine by multiple genome screen. Science 2005, 309, 148–150. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Yılmaz Çolak, Ç. Computational Design of a Multi-epitope Vaccine Against Clostridium chauvoei: An Immunoinformatics Approach. Int. J. Pept. Res. Ther. 2021, 27, 2639–2649. [Google Scholar] [CrossRef]
  117. 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]
  118. Chiang, M.-H.; Sung, W.-C.; Lien, S.-P.; Chen, Y.-Z.; Lo, A.F.; Huang, J.-H.; Kuo, S.-C.; Chong, P. Identification of novel vaccine candidates against Acinetobacter baumannii using reverse vaccinology. Hum. Vaccin. Immunother. 2015, 11, 1065–1073. [Google Scholar] [CrossRef] [Green Version]
  119. Wintachai, P.; Phaonakrop, N.; Roytrakul, S.; Naknaen, A.; Pomwised, R.; Voravuthikunchai, S.P.; Surachat, K.; Smith, D.R. Enhanced antibacterial effect of a novel Friunavirus phage vWU2001 in combination with colistin against carbapenem-resistant Acinetobacter baumannii. Sci. Rep. 2022, 12, 2633. [Google Scholar] [CrossRef]
  120. Ayoub Moubareck, C.; Hammoudi Halat, D. Insights into Acinetobacter baumannii: A review of microbiological, virulence, and resistance traits in a threatening nosocomial pathogen. Antibiotics 2020, 9, 119. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic framework of the methodology used in this study.
Figure 1. Schematic framework of the methodology used in this study.
Ijerph 19 05568 g001
Figure 2. (A). pan- core plot (B). Core phylogeny tree of 50 complete genome of A. baummanni.
Figure 2. (A). pan- core plot (B). Core phylogeny tree of 50 complete genome of A. baummanni.
Ijerph 19 05568 g002
Figure 3. Total, unique, redundant, non-redundant, outer membrane, extracellular and periplasmic membrane, and virulence factors obtained through subtractive proteomics.
Figure 3. Total, unique, redundant, non-redundant, outer membrane, extracellular and periplasmic membrane, and virulence factors obtained through subtractive proteomics.
Ijerph 19 05568 g003
Figure 4. Number of non-antigenic (NA), human homologous (H.H), normal microbiota homologous (N.M.H), allergic and un-stable (U.S) proteins.
Figure 4. Number of non-antigenic (NA), human homologous (H.H), normal microbiota homologous (N.M.H), allergic and un-stable (U.S) proteins.
Ijerph 19 05568 g004
Figure 5. Schematic diagram of 250 amino acids long vaccine construct.
Figure 5. Schematic diagram of 250 amino acids long vaccine construct.
Ijerph 19 05568 g005
Figure 6. 3D structure of vaccine construct. Red color adjuvant, forest green color show GPGPG linkers while yellow color represents epitopes.
Figure 6. 3D structure of vaccine construct. Red color adjuvant, forest green color show GPGPG linkers while yellow color represents epitopes.
Ijerph 19 05568 g006
Figure 7. Original vaccine structure (A) and muted vaccine structure (B). The yellow stick in the mutated vaccine represents disulfide bonds.
Figure 7. Original vaccine structure (A) and muted vaccine structure (B). The yellow stick in the mutated vaccine represents disulfide bonds.
Ijerph 19 05568 g007
Figure 8. Docked intermolecular conformation of the vaccine to (A) MHC-I, (B) MHC-II, and (C) TLR-4.
Figure 8. Docked intermolecular conformation of the vaccine to (A) MHC-I, (B) MHC-II, and (C) TLR-4.
Ijerph 19 05568 g008
Figure 9. Statistical analysis of the molecular dynamics simulation trajectories. RMSD (A), RMSF (B), and RoG (C).
Figure 9. Statistical analysis of the molecular dynamics simulation trajectories. RMSD (A), RMSF (B), and RoG (C).
Ijerph 19 05568 g009
Figure 10. (A) Immunoglobulin titer as shown through different color peaks in response to multi -epitopes vaccine injection as shown with black color peak (B) Elicitation of interleukins level after injection of multi-epitopes vaccine construct as represented by C- immune simulation analysis.
Figure 10. (A) Immunoglobulin titer as shown through different color peaks in response to multi -epitopes vaccine injection as shown with black color peak (B) Elicitation of interleukins level after injection of multi-epitopes vaccine construct as represented by C- immune simulation analysis.
Ijerph 19 05568 g010
Table 1. Physiochemical properties of the selected virulence factor molecules including shortlisted four potential vaccine candidates.
Table 1. Physiochemical properties of the selected virulence factor molecules including shortlisted four potential vaccine candidates.
Outer Membrane ProteinsEncoded ProteinsAmino AcidMolecular WeightGravyAliphatic IndexInstability IndexTheoretical (PI)
>core/109/1/Org1_Gene719Type II toxin-antitoxin system antitoxin maze31071115.65−0.33687.0630.049.35
>core/298/1/Org1_Gene3311Methionine synthase [Mycobacterium tuberculosis77584.63−0.39176.6828.678.58
>core/315/1/Org1_Gene2327Hypothetical protein75880.95−0.22296.3730.488.86
>core/982/1/Org1_Gene2259MULTISPECIES: 1-acyl-sn-glycerol-3-phosphate acyltransferase48452.26−0.31192.0239.599.02
>core/1058/1/Org1_Gene66Nuclear transport factor 2 family protein [Mycobacterium tuberculosis47148.39−0.015100.8132.85.59
>core/2148/1/Org1_Gene1404Hypothetical protein K60_024290 [Mycobacterium tuberculosis variant bovis BCG str. Korea 1168P35638.97−0.39680.93325.32
>core/3446/1/Org1_Gene2677Chromosome segregation protein SMC26630.44−0.41585.2341.449.01
>core/4300/1/Org1_Gene2340NlpC/P60 family peptidoglycan endopeptidase RipB21722.13−0.28884.6130.239.3
>core/1212/7/Org7_Gene3542MULTISPECIES: zf-HC2 domain-containing protein [Mycobacterium tuberculosis complex45149.38−0.4270.35215.66
>core/3840/27/Org27_Gene62MULTISPECIES: hypothetical protein24426.05−0.34780.8242.436.44
>core/4074/16/Org16_Gene452Ribosomal protein S7 from Mycobacterium tuberculosis21125.40.08112.8131.575.74
Extracellular Proteins
>core/331/1/Org1_Gene3212DUF427 domain-containing protein 74283.32−0.55669.4635.456.15
>core/5806/1/Org1_Gene1744D-alanyl-D-alanine carboxypeptidase14315.5-0.0984.0625.085.52
Periplasmic Proteins
>core/3114/1/Org1_Gene3080D-alanyl-D-alanine carboxypeptidase dacb228831.25−0.31895.1419.935.93
>core/466/2/Org2_Gene2561Glycine cleavage system aminomethyltransferase gcvt66169.36−0.24178.1135.085.46
>core/1121/3/Org3_Gene1840MULTISPECIES: multidrug efflux SMR transporter Mmr46349.7−0.11496.235.669.34
>core/2579/16/Org16_Gene524Conserved protein of uncharacterised function32436.57−0.22392.9630.155.81
>core/4275/27/Org27_Gene2284Site-2 protease family protein [Mycobacterium tuberculosis21924.1−0.40176.821.595.88
Table 2. List of potent DRB*0101 binders, probable antigenic, non-allergic, non-toxic, and good water-soluble epitopes.
Table 2. List of potent DRB*0101 binders, probable antigenic, non-allergic, non-toxic, and good water-soluble epitopes.
EpitopesDRB*0101 Binder ScoreAntigenicityAllergenicityWater
Solubility
Toxicity
LEQQQAAQT7.635Probable antigenicNon-AllergicGood water solubleNon-toxic
QDDTYAGGQ12.74
GSKPNRVPI13.55
QGVRNINGR17.02
IYGADGEQR63.53
TNGPELQDD41.78
FQDSQHNNG5.85
TPVAPQPQE52.97
NIKDQYKPE22.86
IADNKTKEG77.62
YVRTVGQKN47.32
YGYRWARDN6.41
KIGENKNAG83.18
NTDADASGR3.29
AADGIAGEE36.73
GYDYGQNAD27.42
NVTDDAQER1.17
Table 3. Top 10 refined docked complexes of vaccine to MHC-1 and model vaccine construct generated by FireDock. ACE (Atomic contact energy) and HB (Hydrogen bonding).
Table 3. Top 10 refined docked complexes of vaccine to MHC-1 and model vaccine construct generated by FireDock. ACE (Atomic contact energy) and HB (Hydrogen bonding).
RankSolution NumberGlobal EnergyAttractive VdWRepulsive VdWACEHB
17−17.35−6.20−1.65−4.99−1.25
2525.32−33.8317.6913.88−0.26
39436.06−52.43675.26−3.37−6.49
411559.57−83.372066.1115.99−8.89
562178.22−50.992740.5929.10−11.89
6104719.15−72.396014.7016.31−8.98
786838.22−94.588700.096.24−14.44
8313,952.77−95.4117,578.4116.24−17.85
9413,994.61−146.0417,861.02−19.01−27.64
10222,692.17−198.3128,778.513.39−32.98
Table 4. Top 10 refined docked complexes of vaccine to MHC-II and model vaccine construct generated by FireDock.
Table 4. Top 10 refined docked complexes of vaccine to MHC-II and model vaccine construct generated by FireDock.
RankSolution NumberGlobal EnergyAttractive VdWRepulsive VdWACEHB
18−2.99−4.291.494.17−1.27
277.20−1.750.001.920.00
3320.83−4.655.591.20−0.27
4923.60−5.650.784.63−0.30
5535.96−25.2536.4114.29−3.90
610164.06−15.38199.7213.40−1.43
741426.58−34.551827.261.63−2.65
862708.43−65.543541.68−9.12−2.25
917877.71−87.439990.6812.16−9.37
10210,558.78−89.7513,391.88−3.44−16.32
Table 5. Top 10 refined docked complexes of vaccine to TLR4 and model vaccine construct generated by FireDock.
Table 5. Top 10 refined docked complexes of vaccine to TLR4 and model vaccine construct generated by FireDock.
RankSolution NumberGlobal EnergyAttractive VdWRepulsive VdWACEHB
12−3.22−26.777.4321.39−2.28
277.57−2.840.001.280.00
3910.66−15.8132.022.23−1.76
41028.75−11.253.6213.23−0.50
5629.86−11.031.7910.45−1.72
61915.94−48.211190.1217.81−4.33
753386.99−52.354311.658.99−5.38
844737.61−121.116190.84−3.73−16.92
937186.19−93.889131.9116.17−11.81
1087763.02−85.239873.192.20−7.86
Table 6. Residues vise interaction of docked complex with MHC-I, MHC-II, and TLR-4 receptors.
Table 6. Residues vise interaction of docked complex with MHC-I, MHC-II, and TLR-4 receptors.
Vaccine
Complex
Interactive Residues
MHC-IAsn346, Asn347, Arg131, Glu58, Glu60, Glu161, Glu53, Ile 46, Lys127, Leu 126, Met138, Pro57, Ser132, Try135, Thr124, Leu64, Phe8, Try50, Try133
MHC-IIIle187, Ser218, Ser240, Val203
TLR-4Asn58, Asp84, Asn176, Asn173, Asn44, Cys 29, Glu42, Arg69, Asn112, Asp84, Arg67, Cys40, Cys148, Cys29, Tyr70, Leu17, Asp35, Thr92, Gln91, Glu136, Gln91, Glu31, Glu179, Glu143, Glu144, Glu111, Glu178, Glu31, His179, His8, Ile138, Tyr46, Glu89, Ile93, Lys108, Pro28, Pro113, Phe64, Phe147, Pro145, Pro142, Phe77, Pro78, Pro78, Leu74, Met40, Ser172, Ser141, Ser141, Ser126, Thr110, Leu85, Lys47, Lys30, Lys130, Tyr46, Leu66, Thr15, Leu152, Lys153, Ser105, Thr106, Lys57, Thr56, Thr37, Tyr38, Thr37, Val133, Thr37, Val35, Val30, Pro28, Val35, Val30.
Table 7. MM-GB/PBSA binding free energies calculation.
Table 7. MM-GB/PBSA binding free energies calculation.
Energy ParameterTLR-4-Vaccine ComplexMHC-I-Vaccine ComplexMHC-II-Vaccine Complex
MM-GBSA
VDWAALS−97.12−106.08−107
EEL−105.68−85.68−88.07
EGB48.0132.3928.24
ESURF−15.04−19.25−15.11
Delta G gas−202.8−191.76−195.07
Delta G solv32.9713.1413.13
Delta Total−169.83−178.62−181.94
MM-PBSA
VDWAALS−97.12−98.36−117.66
EEL−105.68−88.07−54.05
EPB43.252328.28
ENPOLAR−9−14.19−17.1
Delta G gas−202.8−191.76−195.07
Delta G solv34.258.8111.18
Delta Total−168.55−182.95−183.89
VDWAALS (van der Waals), EEL (electrostatic), Delta G gas (net gas phase energy), Delta G solv (net solvation energy), Delta Total (net energy of system).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

ud-din, M.; Albutti, A.; Ullah, A.; Ismail, S.; Ahmad, S.; Naz, A.; Khurram, M.; Haq, M.u.; Afsheen, Z.; Bakri, Y.E.; et al. Vaccinomics to Design a Multi-Epitopes Vaccine for Acinetobacter baumannii. Int. J. Environ. Res. Public Health 2022, 19, 5568. https://doi.org/10.3390/ijerph19095568

AMA Style

ud-din M, Albutti A, Ullah A, Ismail S, Ahmad S, Naz A, Khurram M, Haq Mu, Afsheen Z, Bakri YE, et al. Vaccinomics to Design a Multi-Epitopes Vaccine for Acinetobacter baumannii. International Journal of Environmental Research and Public Health. 2022; 19(9):5568. https://doi.org/10.3390/ijerph19095568

Chicago/Turabian Style

ud-din, Miraj, Aqel Albutti, Asad Ullah, Saba Ismail, Sajjad Ahmad, Anam Naz, Muhammad Khurram, Mahboob ul Haq, Zobia Afsheen, Youness El Bakri, and et al. 2022. "Vaccinomics to Design a Multi-Epitopes Vaccine for Acinetobacter baumannii" International Journal of Environmental Research and Public Health 19, no. 9: 5568. https://doi.org/10.3390/ijerph19095568

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