Skip to main content

Immunoinformatics approach for multi-epitope vaccine design against structural proteins and ORF1a polyprotein of severe acute respiratory syndrome coronavirus-2 (SARS-CoV-2)

Abstract

Background

The lack of effective treatment against the highly infectious SARS-CoV-2 has aggravated the already catastrophic global health issue. Here, in an attempt to design an efficient vaccine, a thorough immunoinformatics approach was followed to predict the most suitable viral proteins epitopes for building that vaccine.

Methods

The amino acid sequences of four structural proteins (S, M, N, E) along with one potentially antigenic accessory protein (ORF1a) of SARS-CoV-2 were inspected for the most appropriate epitopes to be used for building the vaccine construct. Several immunoinformatics tools were used to assess the antigenicity (VaxiJen server), immunogenicity (IEDB immunogenicity tool), allergenicity (AlgPred), toxigenicity (ToxinPred server), interferon-gamma inducing capacity (IFNepitope server), and the physicochemical properties of the construct (ProtParam tool).

Results

The final candidate vaccine construct consisted of 468 amino acids, encompassing 29 epitopes. The CTL epitopes that passed the antigenicity, allergenicity, toxigenicity and immunogenicity assessment were four epitopes from S protein, one from M protein, two from N protein, 12 from the ORF1a polyprotein and none from E protein. While the HTL epitopes that passed the antigenicity, allergenicity, toxigenicity and INF-\(\gamma\) were one from S protein, three from M protein, six from the ORF1a polyprotein and none from N and E proteins.

All the vaccine properties and its ability to trigger the humoral and cell-mediated immune response were validated computationally. Molecular modeling, docking to TLR3, simulation, and molecular dynamics were also carried out. Finally, a molecular clone using pET28::mAID expression plasmid vector was prepared.

Conclusion

The overall results of the study suggest that the final multi-epitope chimeric construct is a potential candidate for an efficient protective vaccine against SARS-CoV-2.

Introduction

In early December 2019, an acute respiratory disease of unknown etiology emerged in Wuhan, China, which was subsequently found to be caused by a novel coronavirus. The virus was initially described as 2019-nCoV and later named by the international committee on taxonomy of viruses (ICTV) as severe acute respiratory syndrome coronavirus-2 (SARS-CoV-2), while the World Health Organization (WHO) named the disease Coronavirus disease-19 (COVID-19) [1,2,3,4,5]. Within the first three months after its discovery, the disease spread to more than 100 countries and caused more than 4,000 deaths worldwide [6]. On the 11th of March, 2020, the WHO categorized the newly discovered disease as a pandemic. COVID-19 is characterized by a broad clinical spectrum, ranging from asymptomatic, to mild to severe respiratory illness requiring intubation and intensive care. The disease course and outcome are contingent on a number of factors, such as age and presence of underlying comorbidities [7]. The clinical manifestations include fever, fatigue, nonproductive cough, dyspnea and myalgia. In severe cases, acute respiratory distress syndrome (ARDS), acute cardiac injury, and acute kidney injury and death can also occur [8, 9].

SARS-CoV-2 along with severe acute respiratory syndrome coronavirus (SARS-CoV) and Middle East respiratory syndrome coronavirus (MERS-CoV) are Betacoronaviruses belonging to the subfamily Othrocoronavirinae of the Nidovirales. These are enveloped, non-segmented, single-stranded, positive-sense RNA viruses, with genomes ranging from 26 to 32 Kb. The genome size of SARS-CoV-2 varies from 29.8 to 29.9 Kb, with typical genome structure of earlier well-characterized coronaviruses, such as the overlapping open reading frame 1a (ORF1a) and 1ab (ORF1ab) region and genes encoding four structural proteins including spike proteins (S), envelope proteins (E), membrane proteins (M), and nucleocapsid proteins (N), in addition to accessory proteins coding genes ORF3a, ORF6, ORF7a, ORF7b and ORF8 [10,11,12,13]. The main role of the spike (S) glycoproteins is to mediate binding to the angiotensin-converting enzyme 2 (ACE-2) receptor and promote membrane fusion and virus entry [14]. Both M and E proteins were reported to play important roles in viral entry, replication, and virions assembly [15]. N proteins are important for viral RNA packaging, virions release and interferon inhibition, promoting the virus pathogenicity [16, 17]. In SARS-CoV, the gene for N protein is upregulated, producing large amounts of the highly immunogenic protein [18]. On the other hand, ORF1a encodes nonstructural polyproteins (PP1a), these polyproteins are involved in viral genome replication and transcription [19].

The COVID-19 pandemic has affected all walks of life, stretching health-care systems to their maximum and putting a huge economical, psychological, and mental burden on the entire world population. This dire situation is aggravated by the contagious nature of the virus, lack of complete understanding of the disease course and the absence of a reliable cure [6]. The disease containment measures used thus far, are contingent on disrupting the transmissibility of the virus through rapid identification and isolation of infected and carrier individuals. This entails the search for vaccines and effective treatments. Recently, number of newly developed vaccines were granted emergency use authorization in many countries worldwide, these are mRNA vaccines such as BNT162b2, and mRNA-1273, DNA vaccines such as AZD1222, Ad26.COV2.S and Sputnik V, inactivated virus vaccines such as CoronaVac and BBIBP-CorV, and protein subunit vaccines such as NVX-CoV2373. Most of these vaccines rely mainly on S protein epitopes, and showed very promising results during different trial phases but they are being closely monitored for any issues regarding their effectiveness and safety. The aim of this study is to design a multi-epitope vaccine against SARS-CoV-2 based on four structural proteins along with the nonstructural polyprotein of ORF1a, using an immunoinformatics approach.

The selection of the nonstructural ORF1a polyprotein alongside the structural viral proteins in this study was driven by suggestions made by a number of studies on other viruses, that nonstructural polyproteins induce immunity and may be applicable to prophylaxis of viral disease [20,21,22,23]. ORF1a was selected over the larger ORF1ab because these two regions overlap, and most of the important proteins found in the region are covered by ORF1a. In addition, ORF1ab is the largest region in viral genome with possibility of larger number of potential epitopes, which in turn may increase the size of the construct to the point that the molecular weight of the final vaccine product will be too large and hinders its effectiveness and delivery.

Materials and methods

Retrieval of target proteins sequences

The amino acid sequences for S protein of 1273 amino acid (Accession No. QLI51913.1), M protein of 222 amino acid (Accession No. QLI52072.1), E protein of 75 amino acid (Accession No. QLI52071.1), N protein of 419 amino acid (Accession No. QIH45060.1) and ORF1a polyprotein of 4405 amino acid (Accession No. QJQ84087.1) were retrieved from NCBI protein database (https://www.ncbi.nlm.nih.gov/protein) in FASTA format.

Cytotoxic T- cell lymphocyte (CTL) epitopes prediction

Initially, the amino acid sequences of all 5 proteins were screened for antigenicity with VaxiJen 2.0 server (http://www.ddg-pharmfac.net/vaxijen/VaxiJen/VaxiJen.html), with threshold value of 0.4 [24]. The CTL epitopes for all sequences were predicted using artificial neural network algorithm-based NetCTL 1.2 server (http://www.cbs.dtu.dk/services/NetCTL/), with threshold value of 0.75 which indicates 0.80 sensitivity and 0.97 specificity [25], which predicts major histocompatibility complex-1 (MHC-1) binding epitopes. The peptides obtained were then checked for antigenicity using VaxiJen 2.0 server. The antigenic peptides were then submitted for virtual scanning for toxic peptides using ToxinPred server (http://crdd.osdd.net/raghava/toxinpred/multi_submit.php), with threshold value 0.0 [26]. The immunogenicity of the resultant non-toxin epitopes was determined using class I immunogenicity tool of Immune Epitope Database (IEDB) (http://tools.iedb.org/immunogenicity/), version 2.22 [27].

Helper T-lymphocyte (HTL) epitopes prediction

For prediction of HTL epitopes, MHC-II binding tool of IEDB (http://tools.iedb.org/mhcii/) was used [28], selecting 7-allele HLA reference set that includes; HLA-DRB1*03:01, HLA-DRB1*07:01, HLA-DRB1*15:01, HLA-DRB3*01:01, HLA-DRB3*02:02, HLA-DRB4*01:01, HLA-DRB5*01:01. The resultant epitopes with low percentile ranks were then checked for allergenicity with AlgPred server (https://webs.iiitd.edu.in/raghava/algpred/submission.html), using the support vector machine (SVM) module based on amino acid composition as the prediction approach [29]. The antigenicity and toxicity status of the non-allergenic epitopes was determined using the VaxiJen 2.0 server and ToxinPred server, respectively. Finally, interferon-gamma (IFN-\(\gamma\)) inducing epitopes were predicted with IFNepitope server (http://crdd.osdd.net/raghava/ifnepitope/) following the Motif and SVM hybrid approach [30]. The resultant epitopes were then inspected for overlapping.

Population coverage

The prediction of worldwide population coverage of the selected epitopes for MHC-I and MHC-II alleles was carried out using population coverage tool of IEDB (http://tools.iedb.org/population/) [31], calculating the coverage for class I and class II separately and combined. The MHC-I alleles assessed included; HLA-B*15:01, HLA-A*30:02, HLA-A*01:01, HLA-B*40:01, HLA-B*07:02, HLA-B*51:01, HLA-A*68:02, HLA-A*02:01, HLA-A*02:06, HLA-B*08:01, HLA-A*02:03, HLA-A*33:01, HLA-A*24:02, HLA-A*23:01, HLA-B*44:03, HLA-B*44:02, HLA-A*31:01, HLA-B*53:01, HLA-A*11:01, HLA-A*68:01, HLA-A*30:01, HLA-B*57:01, HLA-A*03:01, HLA-A*26:01, HLA-B*58:01, HLA-A*32:01, HLA-B*35:0. The world coverage for these alleles was 98.55%. For MHC-II, the alleles assessed included; HLA-DRB1*07:01, HLA-DRB1*15:01, HLA-DRB3*01:01.

B-cell epitopes prediction

The linear B-cell epitopes of all proteins under study were predicted with the antibody epitope prediction tool of IEDB (http://tools.iedb.org/bcell/) using BepiPred linear epitope prediction method 2.0 [32], Emini surface accessibility prediction method [33], and Kolaskar and Tongaonkar antigenicity method [34].

Construction of multiepitope vaccine sequence

To ensure efficient vaccine construction and proper epitope separation, all candidate epitopes were joined together using linkers. The B-cell epitope and CTL epitopes were linked with AAY linker, and HTL epitopes were linked together and to the CTL epitopes with GPGPG linker. To facilitate future conjugation of the multi-epitope vaccine construct with a carrier protein, a cysteine residue was added at the N-terminal [35]. Furthermore, a four amino acid (EPEA) tag was added at the C-terminal for efficient purification [36]. The vaccine construct was subjected to further analysis to assess its antigenicity with VaxiJen 2.0 server, allergenicity with AlgPred server and physicochemical properties with ProtParam tool (https://web.expasy.org/protparam/) [37].

Modeling and structure validation

The secondary structure of the novel vaccine construct was determined using PSIPRED server (http://bioinf.cs.ucl.ac.uk/psipred/) [38]. Protein modeling was carried out using threading and ab initio approaches with IntFOLD and trRosetta servers (https://www.reading.ac.uk/bioinf/IntFOLD/IntFOLD5_form.html) (https://yanglab.nankai.edu.cn/trRosetta/) [39], further protein structure analysis and model validation carried out using ProSA-web server (https://prosa.services.came.sbg.ac.at/prosa.php) [40], Ramachandran plot analysis using RAMPAGE server (http://mordred.bioc.cam.ac.uk/~rapper/rampage.php) [41] and ERRAT server (https://servicesn.mbi.ucla.edu/ERRAT/) [42].

Molecular docking

The vaccine construct was subjected to molecular docking with Toll-like receptor -3 (TLR-3) using FRODOCK (http://frodock.chaconlab.org) and GRAMM-X simulation servers (http://vakser.compbio.ku.edu/resources/gramm/grammx/index_html), with default parameters [43].

The docked vaccine-receptor complex was then prepared for simulation using a protein-prep wizard and PyMOL software using the default settings, the molecular dynamics simulation was then carried out using the Desmond tool and Superpose1.0 server (http://superpose.wishartlab.com) for calculating the root mean square deviation (RMSD).

Immune response simulation

The immune response to the novel multi-epitope vaccine construct was carried out using C-ImmSim server 10.1 (http://www.cbs.dtu.dk/services/C-ImmSim-10.1/) [44]. The simulation parameters used were, random seed: 12,345, simulation steps: 100 and simulation volume: 10 \(\mu\) L. The default injection schedule with the antigen name, injection time: 0 and the injection amount: 1000.

In silico molecular cloning

The amino acid sequence for the candidate vaccine was then subjected to reverse translation and codon optimization with JAVA Codon Adaptation Tool (Jcat) (http://www.jcat.de) [45]. The DNA sequence was then used for in silico molecular cloning with expression plasmid vector pET28::mAID from E.coli [46] using Snapgene software version 5.2.

Results

T-cell epitopes prediction

The initial screening of amino acid sequences of all five proteins for antigenicity, showed a score greater than the threshold value of 0.4 indicating probable antigens, these sequences were then submitted to NetCTL server to predict possible CTL epitopes, which resulted in 37 possible epitopes for S protein, out of which 14 showed no toxicity and eight positive immunogenicity score. Ultimately, the top four epitopes were selected for inclusion in the multi-epitope vaccine construct. For M protein, 10 epitopes were predicted, five epitopes showed no toxicity and only one showed a positive immunogenicity score. For E protein, three epitopes were predicted, two of which showed an antigenicity score higher than the threshold value and non-toxic, but neither showed a positive immunogenicity score, hence, not included in the vaccine construct. For N protein, nine epitopes were predicted, six showed an antigenicity score higher than the threshold value, all six predicted epitopes showed no toxicity, of which, five showed positive immunogenicity score, and only the top two were selected to be included in the construct. For the nonstructural polyprotein, on the other hand, 170 epitopes were predicted, of which 96 showed antigenicity score higher than the threshold, and the best 12 were selected based on toxigenicity and immunogenicity results (Table 1).

Table 1 Cytotoxic T-cell lymphocyte predicted epitopes of selected proteins based on antigenicity, toxicity, and immunogenicity

The HTL epitopes prediction with the MHC-II binding tool of IEDB and based on percentile rank less than 10, resulted in 17 epitopes for S protein, of which 12 were non-allergenic, 10 were non-toxic and a single epitope showed a positive interferon-gamma induction result. For M protein, the predicted HTL epitopes were 55, out of which 43 were non-allergenic antigenic non-toxic epitopes, and only three epitopes showed positive interferon-gamma induction results. None of the predicted HTL epitopes of N protein showed interferon-gamma positive results, therefore none were in the vaccine construct. Similarly, all HTL epitopes predicted for E protein failed to pass either the antigenicity, allergenicity, or interferon-gamma induction assessment. Out of 96 HTL predicted epitopes for ORF1a polyprotein, only six epitopes passed the antigenicity, allergenicity, toxigenicity, and interferon-gamma induction assessment. Results are shown in Table 2.

Table 2 Helper T-cell lymphocyte predicted epitopes of selected proteins based on antigenicity and IFN-γ response

B-cell epitopes prediction

The B-cell epitopes are an important part of the multi-epitope vaccine because recognition of these epitopes by B lymphocytes elicit antibody production, which is a key process in adaptive immunity. For all five proteins, linear B-cell epitopes were predicted using Bepipred Linear Epitope Prediction 2.0 method, Emini surface accessibility prediction method, and Kolaskar & Tongaonkar antigenicity method. These methods were selected because they assess properties that are important for predicting potential epitopes, such as antigenicity, surface accessibility, and flexibility. The resultant plots were then inspected for overlapping regions showing epitopes by the three methods. The only protein to show such an overlapping region was N protein with a sequence of 10 amino acids from 380–390. The results of all amino acid sequences are shown in Fig. 1.

Fig. 1
figure 1

B-cell predicted epitopes of selected proteins using A BepiPred method; B Emini method; C Kolaskar and Tongaonkar method

Population coverage

The selected epitopes were then analyzed to determine the percentage of the world population coverage for MHC-I and MHC-II alleles. The coverage for these alleles was 49.02. The combine allele coverage for both MHC-I and MHC-II was found to be 99.26% which indicates a high population coverage for selected epitopes (Fig. 2).

Fig. 2
figure 2

World population coverage for combined MHC-I and II alleles

Multi-epitope vaccine construction

For the construction of the final vaccine construct, the most appropriate predicted epitopes were selected, this included one B-cell linear epitope from N protein, four CTL and three HTL epitopes from S protein, one CTL and two HTL epitopes from M. protein, two CTL epitopes from N protein, 12 CTL and six HTL epitopes from ORF1a. These epitopes were joined together with two types of linkers, AAY for linear B-cell and CTL epitopes, and GPGPG for HTL epitopes, with cysteine residue at the N-terminal and EPEA tag at C-terminal, this yielded the following 468 amino acid peptide chain:

CQALPQRQKKQQAAYQLTPTWRVYAAYVLPFNDGVYAAYWTAGAAAYYAAYCNDPFLGVYAAYAGDSGFAAYAAYLSPRWYFYYAAYSPDDQIGYYAAYVSDIDITFLAAYTLRVEAFEYAAYHVGEIPVAYAAYSTNVTIATYAAYLVSDIDITFAANGDVVAIDYAAYVVDYGARFYAAYGTDPYEDFQAAYVTNNTFTLKAAYETSWQTGDFAAYFMGRIRSVYAAYVVVNAANVYGPGPGTRFASVYAWNRKRISGPGPGFQTLLALHRSYLTPGGPGPGQPYRVVVLSFELLHAGPGPGSRTLSYYKLGASQRVGPGPGLVGLMWLSYFIASFRGPGPGVSTQEFRYMNSQGLLGPGPGAALGVLMSNLGMPSYGPGPGTLNGLWLDDVVYCPRGPGPGAYESLRPDTRYVLMDGPGPGSAGIFGADPIHSLRVGPGPMFTPLVPFWITIAYIGPGPGEPEA.

Physiochemical properties of the vaccine construct

The results obtained from the ProtParam server, showed that the novel vaccine construct has a molecular weight of 50.417 KDa which is an optimum molecular weight for an antigenic protein. The theoretical isoelectric point (PI) for the construct was 5.41 indicating an acidic nature, with a total of 30 negatively charged residues and 25 positively charged residues. The estimated half-life is 1.2 h in mammalian reticulocytes in vitro, > 20 h in yeast in vivo, and > 10 h in E. coli in vivo, indicating a good construct for future cloning. The instability index was computed to be 30.79 suggesting stable protein. The aliphatic index of 75.38, which indicates a thermostable protein. The grand average of hydropathicity (GRAVY) was 0.040, a positive value close to zero means a slightly hydrophobic molecule.

Vaccine modeling and structure analysis

Based on the amino acid sequence of the vaccine construct, the result of the PSIPRED server revealed different secondary structures (Fig. 3). This is considered a primary step towards predicting the three-dimensional structure of the protein.

Fig. 3
figure 3

Secondary structure prediction of the novel vaccine construct

The 3D protein model was then predicted with two modeling approaches; threading model with IntFOLD server and ab initio modeling with the trRosetta server, the resultant models were then analyzed with Ramachandran plot and ProSA-web z-score based on X-ray crystallography and NMR analysis. The best-predicted model showed 98% of the residues in the favorable region in Ramachandran plot (Fig. 4A), and z-score of − 6.01, determined by x-ray crystallography (Fig. 4B).

Fig. 4
figure 4

A Ramachandran plot showing 98% residues in the favorable region, B z-score determined by x-ray crystallography showing value of − 6.01

The statistics of non-bonded interactions between different atom types and the error function value was plotted against a position of a 9-residue sliding window, calculated by comparison with statistics from highly refined structures, carried out using ERRAT server, and the calculated error value obtained was 81.928, which falls well below 91% indicating a relatively average overall quality for the selected protein model. This is can be justified by the fact that the modeling process was carried out using ab initio modeling approach (Fig. 5A).

Fig. 5
figure 5

A 3D structure model of the candidate vaccine, B Docking of the vaccine construct with Toll-like receptor-3

Molecular docking and dynamics

The final vaccine construct was docked with Toll-like receptor 3 (PDB ID: 1ziw) using the FRODOCK server. The value of the root mean square deviation was 3.78 which suggests a relatively poor binding pose at the site of the receptor and vaccine binding Fig. 5B.

Immune response simulation

Measuring the immune response is a pivotal step for vaccine designing and this is contingent on a number of algorithms that make use of mathematical models to illustrate the fine details of the immunological process. In the present study, the C-ImmSim server was used to simulate immune response with the candidate vaccine construct. Simulation with this tool focuses on B-cell epitope binding, class I and II HLA epitope binding, and the binding of the T-cell receptor to HLA-peptide complexes [47].

The simulation results showed an increased and sustained level of B- memory and active cells, and a high level of IgM, which represents the primary response against the antigen and this suggests effective humoral response (Fig. 6A, B). T helper cell population showed very promising results, as the levels of memory helper cells and active T helper cells remained high for the entire period of simulation, suggesting prolonged humoral and cell-mediated immune response (Fig. 6C, D). The results of the T cytotoxic cell population showed a steady level of the memory cells, while the active cell population showed an increased level throughout the stimulation period (Fig. 6E, F). The result of different immunoglobulin isotopes showed high level in the first two weeks followed by a gradual decline, similar result was shown by interferon-gamma level. This can be viewed as a positive point, since, the first two weeks are considered detrimental for the course and outcome of the disease (Fig. 6G, H) [48].

Fig. 6
figure 6

Different immune responses simulation of the vaccine construct using C-ImmSim. A B Cell population (cells/mm3), B B Cell population per state (cells/mm3), C TH Cell population (cells/mm3), D TH Cell population (cells/mm3), E TC cell population (cells/mm3), F TC cell population per state (cell/mm3), G Concentration of immunoglobulins & immunocomplexes, H Concentration of cytokines & interleukins

In silico molecular cloning

The DNA sequence produced by Jcat showed a GC content of 56% and a codon adaptation index of 1.0, which indicate a stable DNA sequence and a high level of protein expression (Fig. 7).

Fig. 7
figure 7

In silico cloning of vaccine construct using pET28b plasmid

Discussion

The current COVID-19 pandemic associated with SARS-CoV-2 infection is the third coronavirus outbreak in the last 20 years besides the severe acute respiratory syndrome (SARS) and the Middle East respiratory syndrome (MERS). SARS-CoV-2 shows relatively higher transmissibility as compared to other emerging viruses such as H7N9 and MERS-CoV [49, 50]. This entails the imperative search for effective vaccine and treatment in addition to the protective and social distancing measures to contain and control the disease. The immunoinformatics approach provides a promising tool for designing and exploring potential vaccines against bacterial, parasitic, and viral diseases [51]. In this study, a multi-epitope vaccine was constructed using the virus structural proteins and the largest non-structural polyprotein [52]. These proteins were selected based on suggestions from previous studies [53, 54]. Unlike the single subunit vaccine, the multi-epitope vaccine is believed to induce a better and more protective immune response [55].

A number of previously conducted studies used similar approach to construct multi-epitope vaccines, however, unlike our present study, these previous attempts used either the structural proteins alone for constructing the vaccine [56, 57] or the spike protein and one non-structural protein [58], or entire set of viral proteins.

In the present study antigenic, non-allergenic, and non-toxic epitopes were identified and used for the construction of the final candidate vaccine. All five proteins were studied for potential epitopes, however, none of the peptides from the envelope protein (E) was eligible for the selection in the final vaccine construct, due to either lack of antigenicity or the allergenicity and toxicity of these peptides, this can also be attributed to the small size of the protein. For the final vaccine construct, CTL, HTL, line B-cell epitopes were linked together using AAY, and GPGPG linkers which provide proper proteasomes cleavage sites for different immune cells [59] which will ultimately enhance the antigen presentation process by binding transporters associated with antigen processing (TAP) [60]. Furthermore, linking of CTL epitopes from different proteins together forms epitopes on a string which is believed to enhance the immunogenicity of CTL epitopes [61]. To the N-terminal of the vaccine construct a cysteine residue was added to facilitate the binding of this vaccine to protein carrier [35], and to the C-terminal, a small peptide of four amino acids EPEA was added to enable downstream purification process [36]. The candidate vaccine construct consists of 486 amino acids, which is an ideal vaccine length, since larger proteins are presented by dendritic cells leading to stronger T-cell immune response [60], while extremely short peptides may induce tolerance and energy by directly binding MHC molecules of non-professional antigen-presenting cells [62]. Determination of the secondary structure of the protein is a pivotal step towards the prediction of its three-dimensional structure, therefore, the secondary structure of the candidate vaccine was determined using PSIPRED server, followed by structure refinement, and protein modeling. Two approaches were used for modeling the protein, threading approach, and ab initio approach, the best resultant model was selected based on the Ramachandran plot and z-score analyses. The docking of the vaccine and TLR-3 showed a possible hydrophilic interaction [63], this interaction indicates a possible recognition of the vaccine by APC specific receptor, which in turn promotes the immune response [64]. The results of immune response simulation showed very promising results, with a sustained response for the cells involved in the humoral and cell-mediated immunity against SARS-CoV-2. Even though most of the currently in-use vaccines are showing high degrees of effectiveness and safety, the potential future risks cannot be overlooked. Three of these vaccines elicit the immune response against a single viral protein, namely, S protein, however, the recent emergence of number of new variants has cast doubt on the effectiveness of the currently used vaccines, with reports claim that the E484K mutation found in the South African (B.1.351) and Brazilian (B.1.128) variants has a negative impact on the longevity of the neutralizing antibodies and, possibly, the vaccine effectiveness [65]. Other studies reported reduced protection of BNT162b2 vaccine against B.1.351 variant [66], and lack of protection of ChAdOx1 nCoV-19 vaccine against the same variant [67]. Many of the recently identified mutations occur in the viral spike gene conferring antibody neutralization resistance [68], and the accumulation of such mutations is believed to ultimately render the current vaccines directed against the viral spike protein ineffective [69]. The proposed multi-epitope vaccine is designed using several structural and non-structural proteins which makes it an appropriate alternative.

The mRNA vaccines also require certain environmental conditions for preservation of the highly unstable nucleic acid. On the other hand, the inactivated vaccines, though elicit a comprehensive immune response against larger number of viral proteins, there are inherent problems associated with viral inactivation process, and the time-consuming production process.

The conventional methods of vaccine development are very costly and time-consuming, alternatively, the immunoinformatics approach has attracted the attention as an ideal method for designing less-expensive, rapid, efficient, multi-epitope vaccines. However, experimental validation is of utmost importance to ensure the safety and efficacy of the resultant vaccine, it is also beyond the scope of this study to explore any possible pathogenic priming or autoimmune disease induction of the proposed vaccine.

Conclusion

The highly contagious nature of SARS-CoV-2 left the entire world population with no option but to wait for the production of a safe and protective vaccine to break the chain of infection and tackle the spread of this pandemic. It is rather impractical to rely on the conventional methods for producing such a vaccine due to a number of limiting factors. This study is an attempt to design an efficient multi-epitope chimeric subunit vaccine that is capable of mounting a strong immune response by induction of both humoral and cellular mediated immunity, with the help of a large number of immunoinformatics tools. The vaccine construct effectively fulfilled the requirements for characteristics such as antigenicity, allergenicity, immunogenicity, physiochemical properties, eliciting the immune response in a simulation model. It is concluded that this novel construct represents a promising candidate for an efficient protective vaccine against SARS-CoV-2.

Availability of data and materials

Not applicable.

Abbreviations

ACE-2:

Angiotensin converting enzyme 2

ARDS:

Acute respiratory distress syndrome

Covid-19:

Coronavirus disease-19

CTL:

Cytotoxic T lymphocyte

DNA:

Deoxyribonucleic acid

E. protein:

Envelope protein

GRAVY:

Grand average of hydropathicity

HLA:

Human leukocyte antigen

HTL:

Helper T lymphocyte

ICTV:

International committee on Taxonomy of Viruses

IEDB:

Immune epitope database

IFN-\(\gamma\) :

Interferon gamma

Kb:

Kelo base

KDa:

Kelo Dalton

M. protein:

Membrane protein

MERS-CoV:

Middle East Respiratory Syndrome coronavirus

MHC:

Major histocompatibility complex

mRNA:

Messenger RNA

N. protein:

Nucleocapsid protein

NMR:

Nuclear magnetic resonance

ORF:

Open reading frame

PDB:

Protein database

PI:

Isoelectric point

PP1a:

Polyproteins 1a

RMSD:

Root mean square deviation

RNA:

Ribonucleic acid

S. protein:

Spike protein

SARS-CoV-2:

Sever acute respiratory syndrome coronavirus-2

SVM:

Support vector machine

TLR:

Toll-like receptor

WHO:

World Health Organization

References

  1. Fagbule OF. Novel coronavirus. Ann Ib Postgrad Med. 2019;2019(17):108–10.

    Google Scholar 

  2. Du Toit A. Outbreak of a novel coronavirus. Nat Rev Microbiol. 2020;18:123.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  3. Wang C, Horby PW, Hayden FG, Gao GF. A novel coronavirus outbreak of global health concern. Lancet. 2020;395:470–3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Xu YH, Dong JH, An WM, Lv XY, Yin XP, Zhang JZ, et al. Clinical and computed tomographic imaging features of novel coronavirus pneumonia caused by SARS-CoV-2. J Infect. 2020;80:394–400.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Wu YC, Chen CS, Chan YJ. The outbreak of COVID-19: an overview. J Chin Med Assoc. 2020;83:217–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Park SE. Epidemiology, virology, and clinical features of severe acute respiratory syndrome-coronavirus-2 (SARS-CoV-2; Coronavirus Disease-19). Clin Exp Pediatr. 2020;63:119–24.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Sun Y, Koh V, Marimuthu K, Ng OT, Young B, Vasoo S, et al. Epidemiological and clinical predictors of COVID-19. Clin Infect Dis. 2020;71:786–92.

    Article  CAS  PubMed  Google Scholar 

  8. Wang D, Hu B, Hu C, Zhu F, Liu X, Zhang J, et al. Clinical characteristics of 138 hospitalized patients with 2019 novel coronavirus–infected pneumonia in Wuhan, China. JAMA. 2020;323:1061–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Huang C, Wang Y, Li X, Ren L, Zhao J, Hu Y, et al. Clinical features of patients infected with 2019 novel coronavirus in Wuhan, China. Lancet. 2020;395:497–506.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Khailany RA, Safdar M, Ozaslan M. Genomic characterization of a novel SARS-CoV-2. Gene Reports. 2020;19:100682.

    Article  PubMed  PubMed Central  Google Scholar 

  11. Astuti I, Yasrafil A. Severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2): an overview of viral structure and host response. Diabetes Metabol Syndr. 2020;14:407–4012.

    Article  Google Scholar 

  12. Zhang G, Zhang J, Wang B, Zhu X, Wang Q, Qiu S. Analysis of clinical characteristics and laboratory findings of 95 cases of 2019 novel coronavirus pneumonia in Wuhan, China: a retrospective analysis. Respir Res. 2020;21:s12931.

    Article  Google Scholar 

  13. Phan T. Novel coronavirus: from discovery to clinical diagnostics. Infect Genet Evol. 2020;79:104211.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Ou X, Liu Y, Lei X, Li P, Mi D, Ren L, et al. Characterization of spike glycoprotein of SARS-CoV-2 on virus entry and its immune cross-reactivity with SARS-CoV. Nat Commun. 2020;11:1–12.

    Article  Google Scholar 

  15. Bianchi M, Benvenuto D, Giovanetti M, Angeletti S, Ciccozzi M, Pascarella S. Sars-CoV-2 envelope and membrane proteins: structural differences linked to virus characteristics? BioMed Res Int. 2020;2020:78.

    Article  CAS  Google Scholar 

  16. Zeng W, Liu G, Ma H, Zhao D, Yang Y, Liu M, Mohammed A, Zhao C, Yang Y, Xie J, Ding C. Biochemical characterization of SARS-CoV-2 nucleocapsid protein. Biochem Biophys Res Commun. 2020;527:618–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Alanagreh LA, Alzoughool F, Atoum M. The human coronavirus disease COVID-19: its origin, characteristics, and insights into potential drugs and its mechanisms. Pathogens. 2020;9:331.

    Article  CAS  PubMed Central  Google Scholar 

  18. Cong Y, Ulasli M, Schepers H, Mauthe M, V’kovski P, Kriegenburg F, et al. Nucleocapsid protein recruitment to replication-transcription complexes plays a crucial role in coronaviral life cycle. J Virol. 2020;94:e01925.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Chen Y, Liu Q, Guo D. Emerging coronaviruses: genome structure, replication, and pathogenesis. J Med Virol. 2020;92:418–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Gibson CA, Schlesinger JJ, Barrett AD. Prospects for a virus non-structural protein as a subunit vaccine. Vaccine. 1988;6:7–9.

    Article  CAS  PubMed  Google Scholar 

  21. Ip PP, Boerma A, Regts J, Meijerhof T, Wilschut J, Nijman HW, Daemen T. Alphavirus-based vaccines encoding nonstructural proteins of hepatitis C virus induce robust and protective T-cell responses. Mol Ther. 2014;22:881–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Ludert JE, Reyes-Sandoval A. The dual role of the antibody response against the flavivirus non-structural protein 1 (NS1) in protection and immunopathogenesis. Front Immunol. 2019;10:1651.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  23. Henriques HR, Rampazo EV, Gonçalves AJ, Vicentin EC, Amorim JH, Panatieri RH, et al. Targeting the non-structural protein 1 from dengue virus to a dendritic cell population confers protective immunity to lethal virus challenge. PLoS Negl Trop Dis. 2013;7:e2330.

    Article  PubMed  PubMed Central  Google Scholar 

  24. Irini A, Doytchinova S, Darren RF. VaxiJen: a server for prediction of protective antigens, tumour antigens and subunit vaccines. BMC Bioinf. 2007;8:4.

    Article  CAS  Google Scholar 

  25. Larsen MV, Lundegaard C, Lamberth K, Buus S, Lund O, Nielsen M. Large-scale validation of methods for cytotoxic T-lymphocyte epitope prediction. BMC Bioinf. 2007;8:424.

    Article  CAS  Google Scholar 

  26. Gupta C, et al. In silico approach for predicting toxicity of peptides and proteins. PLoS ONE. 2013;8(9):e73957.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Calis JJA, Maybeno M, Greenbaum JA, Weiskopf D, De Silva AD, Sette A, et al. Properties of MHC class I presented peptides that enhance immunogenicity. PloS Comp Biol. 2013;9:e1003266.

    Article  Google Scholar 

  28. Jensen KK, Andreatta M, Marcatili P, Buus S, Greenbaum JA, Yan Z, et al. Improved methods for predicting peptide binding affinity to MHC class II molecules. Immunology. 2018;154:394–406.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Saha S, Raghava GP. AlgPred: prediction of allergenic proteins and mapping of IgE epitopes. Nucleic Acids Res. 2006;34:202–9.

    Article  CAS  Google Scholar 

  30. Dhanda SK, Vir P, Raghava GP. Designing of interferon-gamma inducing MHC class-II binders. Biol Direct. 2013;8:30.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  31. Bui HH, Sidney J, Dinh K, Southwood S, Newman MJ, Sette A. Predicting population coverage of T-cell epitope-based diagnostics and vaccines. BMC Bioinf. 2006;7:1–5.

    Article  CAS  Google Scholar 

  32. Larsen JE, Lund O, Nielsen M. Improved method for predicting linear B-cell epitopes. Immunome Res. 2006;2:1–7.

    Article  CAS  Google Scholar 

  33. Emini EA, Hughes JV, Perlow D, Boger J. Induction of hepatitis A virus-neutralizing antibody by a virus-specific synthetic peptide. J Virol. 1985;55:836–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Kolaskar AS, Tongaonkar PC. A semi-empirical method for prediction of antigenic determinants on protein antigens. FEBS Lett. 1990;276:172–4.

    Article  CAS  PubMed  Google Scholar 

  35. Bandyopadhyay A, Cambray S, Gao J. Fast and selective labeling of N-terminal cysteines at neutral pH via thiazolidino boronate formation. Chem Sci. 2016;7:4589–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Jin J, Hjerrild KA, Silk SE, Brown RE, Labbé GM, Marshall JM, et al. Accelerating the clinical development of protein-based vaccines for malaria by efficient purification using a four amino acid C-terminal ‘C-tag.’ Int J Parasitol. 2017;47:435–46.

  37. Gasteiger E, Hoogland C, Gattiker A, Wilkins MR, Appel RD, Bairoch A. Protein identification and analysis tools on the ExPASy server. The proteomics protocols handbook. Totowa: Humana press; 2005. p. 571–607.

    Book  Google Scholar 

  38. Jones DT. Protein secondary structure prediction based on position-specific scoring matrices. J Mol Biol. 1999;292:195–202.

    Article  CAS  PubMed  Google Scholar 

  39. Yang J, Anishchenko I, Park H, Peng Z, Ovchinnikov S, Baker D. Improved protein structure prediction using predicted interresidue orientations. Proc Natl Acad Sci. 2020;117:1496–503.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Sippl MJ. Recognition of errors in three-dimensional structures of proteins. Proteins. 1993;17:355–62.

    Article  CAS  PubMed  Google Scholar 

  41. Prisant MG, Richardson JS, Richardson DC. Structure validation by Calpha geometry: Phi, psi and Cbeta deviation. Proteins. 2003;50:437–50.

    Article  PubMed  CAS  Google Scholar 

  42. Colovos C, Yeates TO. Verification of protein structures: patterns of nonbonded atomic interactions. Protein Sci. 1993;2:1511–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Tovchigrechko A, Vakser IA. Development and testing of an automated approach to protein docking. Proteins. 2005;60:296–301.

    Article  CAS  PubMed  Google Scholar 

  44. 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.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  45. Grote A, Hiller K, Scheer M, Münch R, Nörtemann B, Hempel DC, Jahn D. JCat: a novel tool to adapt codon usage of a target gene to its potential expression host. Nucleic Acids Res. 2005;33:W526–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Besmer E, Market E, Papavasiliou FN. The transcription elongation complex directs activation-induced cytidine deaminase-mediated DNA deamination. Mol Cell Biol. 2006;26:4378–85.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Rapin N, Lund O, Castiglione F. Immune system simulation online. Bioinformatics. 2011;27:2013–4.

    Article  CAS  PubMed  Google Scholar 

  48. Bar-On YM, Flamholz A, Phillips R, Milo R. Science Forum: SARS-CoV-2 (COVID-19) by the numbers. Elife. 2020;9:e57309.

    Article  PubMed  PubMed Central  Google Scholar 

  49. Dhama K, Khan S, Tiwari R, Sircar S, Bhat S, Malik YS, et al. Coronavirus Disease 2019 – COVID-19. Clin Microbiol Rev. 2020;33:e00028-e120.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Tobaiqy M, Qashqary M, Al-Dahery S, Mujallad A, Hershan AA, Kamal MA, et al. Therapeutic management of patients with COVID-19: a systematic review. Infect Prev Pract. 2020;2:100061.

    Article  PubMed  PubMed Central  Google Scholar 

  51. Kaur R, Arora N, Jamakhani MA, Malik S, Kumar P, Anjum F, et al. Development of multi-epitope chimeric vaccine against Taenia solium by exploring its proteome: an in silico approach. Expert Rev Vaccines. 2020;19:105–14.

    Article  CAS  PubMed  Google Scholar 

  52. Amawi H, Abu Deiab GA, Aljabali AA, Dua K, Tambuwala MM. COVID-19 pandemic: an overview of epidemiology, pathogenesis, diagnostics and potential vaccines and therapeutics. Therapeutic Delivery. 2020;11:245–68.

    Article  CAS  PubMed  Google Scholar 

  53. Ahmed SF, Quadeer AA, McKay MR. Preliminary identification of potential vaccine targets for the COVID-19 coronavirus (SARS-CoV-2) based on SARS-CoV immunological studies. Viruses. 2020;12:254.

    Article  CAS  PubMed Central  Google Scholar 

  54. Wang J, Wen J, Li J, Yin J, Zhu Q, Wang H, et al. Assessment of immunoreactive synthetic peptides from the structural proteins of severe acute respiratory syndrome coronavirus. Clin Chem. 2003;49:1989–96.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Saadi M, Karkhah A, Nouri HR. Development of a multi-epitope peptide vaccine inducing robust T cell responses against brucellosis using immunoinformatics based approaches. Infect Genet Evol. 2017;51:227–34.

    Article  CAS  PubMed  Google Scholar 

  56. Singh A, Thakur M, Sharma LK, Chandra K. Designing a multi-epitope peptide based vaccine against SARS-CoV-2. Sci Rep. 2020;10(1):1–2.

    Article  CAS  Google Scholar 

  57. Dong R, Chu Z, Yu F, Zha Y. Contriving multi-epitope subunit of vaccine for COVID-19: immunoinformatics approaches. Front Immunol. 2020;28(11):1784.

    Article  CAS  Google Scholar 

  58. Safavi A, Kefayat A, Mahdevar E, Abiri A, Ghahremani F. Exploring the out of sight antigens of SARS-CoV-2 to design a candidate multi-epitope vaccine by utilizing immunoinformatics approaches. Vaccine. 2020;38(48):7612–28.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Chen X, Zaro JL, Shen WC. Fusion protein linkers: property, design and functionality. Adv Drug Deliv Rev. 2013;65:1357–69.

    Article  CAS  PubMed  Google Scholar 

  60. Nezafat N, Ghasemi Y, Javadi G, Khoshnoud MJ, Omidinia E. A novel multi-epitope peptide vaccine against cancer: an in silico approach. J Theor Biol. 2014;349:121–34.

    Article  CAS  PubMed  Google Scholar 

  61. Livingston B, Crimi C, Newman M, Higashimoto Y, Appella E, Sidney J, et al. A rational strategy to design multiepitope immunogens based on multiple Th lymphocyte epitopes. J Immunol. 2002;168:5499–506.

    Article  CAS  PubMed  Google Scholar 

  62. Melief CJ, Van Der Burg SH. Immunotherapy of established (pre) malignant disease by synthetic long peptide vaccines. Nat Rev Cancer. 2008;8:351–60.

    Article  CAS  PubMed  Google Scholar 

  63. Botos I, Segal DM, Davies DR. The structural biology of Toll-like receptors. Structure. 2011;19:447–59.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Droppa-Almeida D, Franceschi E, Padilha FF. Immune-informatic analysis and design of peptide vaccine from multi-epitopes against Corynebacterium pseudotuberculosis. Bioinform Biol Insights. 2018;12:1177932218755337.

    Article  PubMed  PubMed Central  Google Scholar 

  65. Wise J. Covid-19: The E484K mutation and the risks it poses. BMJ. 2021;372:n359.

    Article  PubMed  Google Scholar 

  66. Abu-Raddad LJ, Chemaitelly H, Butt AA; National Study Group for COVID-19 Vaccination. Effectiveness of the BNT162b2 Covid-19 Vaccine against the B1.1.7 and B.1.351 Variants. N Engl J Med. 2021. Doi: https://doi.org/10.1056/NEJMc2104974.

  67. Madhi SA, Baillie V, Cutland CL, Voysey M, Koen AL, Fairlie L, Padayachee SD, Dheda K, Barnabas SL, Bhorat QE, Briner C. Efficacy of the ChAdOx1 nCoV-19 Covid-19 vaccine against the B. 1.351 variant. N Engl J Med. 2021. https://doi.org/10.1056/NEJMoa2102214.

    Article  PubMed  PubMed Central  Google Scholar 

  68. McCarthy KR, Rennick LJ, Nambulli S, Robinson-McCarthy LR, Bain WG, Haidar G, Duprex WP. Recurrent deletions in the SARS-CoV-2 spike glycoprotein drive antibody escape. Science. 2021;371(6534):1139–42.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Wang P, Nair MS, Liu L, Iketani S, Luo Y, Guo Y, Wang M, Yu J, Zhang B, Kwong PD, Graham BS. Antibody resistance of SARS-CoV-2 variants B. 1.351 and B. 1.1. 7. Nature. 2021;8:1–6.

Download references

Acknowledgements

I am indebted to the faculty members of college of applied medical sciences and the deanship of scientific research, University of Bisha.

Funding

Deanship of scientific research at University of Bisha, COVID-19 initiative project, Grant No. (UB-COVID-15-1441).

Author information

Authors and Affiliations

Authors

Contributions

The author read and approved the final manuscript.

Corresponding author

Correspondence to Khalid Mohamed Adam.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The author declare that he has no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Adam, K.M. Immunoinformatics approach for multi-epitope vaccine design against structural proteins and ORF1a polyprotein of severe acute respiratory syndrome coronavirus-2 (SARS-CoV-2). Trop Dis Travel Med Vaccines 7, 22 (2021). https://doi.org/10.1186/s40794-021-00147-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s40794-021-00147-1

Keywords