Next Article in Journal
Biodegradable PLGA-b-PEG Nanoparticles Induce T Helper 2 (Th2) Immune Responses and Sustained Antibody Titers via TLR9 Stimulation
Previous Article in Journal
Intranasal Therapeutic Peptide Vaccine Promotes Efficient Induction and Trafficking of Cytotoxic T Cell Response for the Clearance of HPV Vaginal Tumors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Specificity of CD8+ T-Cell Responses Following Vaccination with Conserved Regions of HIV-1 in Nairobi, Kenya

1
The Jenner Institute, University of Oxford, Oxford OX3 7DQ, UK
2
Department of Microbiology and Immunology, Faculty of Pharmacy, Al-Azhar University, Cairo 11823, Egypt
3
Joint Research Center for Human Retrovirus Infection, Kumamoto University, Kumamoto 860-0811, Japan
4
International AIDS Vaccine Initiative IAVI-Human Immunology Laboratory, Imperial College London, London SW10 9NH, UK
5
International AIDS Vaccine Initiative-New York, New York, NY 10004, USA
6
KAVI-Institute of Clinical Research, University of Nairobi, Nairobi 19676 00202, Kenya
7
IrsiCaixa AIDS Research Institute-HIVACAT, Hospital Universitari Germans Trias i Pujol, 08916 Barcelona, Spain
8
Faculty of Medicine, Universitat de Vic-Central de Catalunya (UVic-UCC), 08500 Vic, Spain
9
Catalan Institution for Research and Advanced Studies (ICREA), 08010 Barcelona, Spain
*
Author to whom correspondence should be addressed.
Vaccines 2020, 8(2), 260; https://doi.org/10.3390/vaccines8020260
Submission received: 4 May 2020 / Revised: 20 May 2020 / Accepted: 25 May 2020 / Published: 29 May 2020
(This article belongs to the Section HIV Vaccines)

Abstract

:
Sub-Saharan Africa carries the biggest burden of the human immunodeficiency virus type 1 (HIV-1)/AIDS epidemic and is in an urgent need of an effective vaccine. CD8+ T cells are an important component of the host immune response to HIV-1 and may need to be harnessed if a vaccine is to be effective. CD8+ T cells recognize human leukocyte antigen (HLA)-associated viral epitopes and the HLA alleles vary significantly among different ethnic groups. It follows that definition of HIV-1-derived peptides recognized by CD8+ T cells in the geographically relevant regions will critically guide vaccine development. Here, we study fine details of CD8+ T-cell responses elicited in HIV-1/2-uninfected individuals in Nairobi, Kenya, who received a candidate vaccine delivering conserved regions of HIV-1 proteins called HIVconsv. Using 10-day cell lines established by in vitro peptide restimulation of cryopreserved PBMC and stably HLA-transfected 721.221/C1R cell lines, we confirm experimentally many already defined epitopes, for a number of epitopes we define the restricting HLA molecule(s) and describe four novel HLA-epitope pairs. We also identify specific dominance patterns, a promiscuous T-cell epitope and a rescue of suboptimal T-cell epitope induction in vivo by its functional variant, which all together inform vaccine design.

1. Introduction

Human leukocyte antigen (HLA) molecules are highly polymorphic cell surface glycoproteins, which bind and present processed antigenic peptides derived from viruses, such as human immunodeficiency virus type 1 (HIV-1) to the cells of the immune system [1,2]. Cytotoxic T lymphocytes (CTL) recognize peptide-loaded HLA class I through their T-cell receptor (TCR) and trigger apoptotic pathways in virus-infected cells serving as virus factories, thereby limit production of new infectious virions [3,4]. Polymorphisms in HLA class I alleles are generally clustered within the sequences encoding the HLA peptide-binding groove, which selects for a range of bound peptide specificities that determine the successfully processed, host-cell presented peptidome. This finite set of epitopes combined with the host’s TCR repertoire largely determines whether or not the host can mount an effective T-cell response [5,6]. Viruses with highly variable genomes such as HIV-1 rapidly mutate epitopes to escape T-cell recognition and the immune pressure selects the fittest escaped variants to overgrow [7,8,9,10]. This immediately suggests that epitopes that are easily escaped with minimal fitness cost are less protective than epitopes within functionally constraint, and therefore conserved protein regions [11,12,13,14,15]. HLA haplotype, epitope specificity of the immune responses and escape in targeted epitopes are among the major factors determining the rate of disease progression in untreated HIV-1 infection [16,17,18,19]. Despite the extensive HLA polymorphism and based on peptide binding preferences, many HLA class I alleles can be clustered into 9 supertypes with common structural features and overlapping associated peptidomes [18,20,21,22]. On the population level, common HLAs/HLA supertypes are theoretically less protective relative to the rare ones due to the increased probability of transmission of already escaped viruses; however, through selection of resistant individuals, rare HLAs will eventually become more common and, therefore, less protective [12,22,23,24,25,26]. Thus, hosts with their HLA haplotypes and associated peptidomes are constantly shaped by much faster evolving infectious agents and other diseases.
T cells clearly play a protective role in HIV-1 infection [27,28,29,30,31] and HIV-1 T-cell vaccine development would be greatly facilitated by a definition of functional correlates of T-cell protection [32,33,34,35,36] including cumulative identification of protective epitopes in vulnerable HIV-1 sites [37,38,39,40,41,42,43]. The current strategies for vaccine development are likely biased towards inclusion of immunodominant epitopes restricted by most frequent/studied HLA class I alleles, and HLAs associated with a slow disease progression [14,44,45,46,47,48,49]. In contrast, our working hypothesis postulates that focusing vaccine-elicited T cells on the functionally conserved regions of HIV-1, which are common to most global variants and are hard to mutate, will be effective in slowing and controlling HIV-1 infection [41,42,43,50,51,52,53,54,55]. These regions contain epitopes typically subdominant in natural HIV-1 infection, but are capable of inducting robust T-cell responses when delivered by a potent vaccination regimen [50,56,57,58,59,60]. Such vaccine-elicited responses thus represent a rich source of previously undescribed, potentially important epitopes [61,62,63]. In the present work, we characterize CD8+ T-cell responses induced by the first generation of conserved-region vaccines expressing immunogen HIVconsv [53] in an HIV-1-negative population in the Kangemi district of Nairobi, Kenya [60], and confirm a number of already-known as well as identify novel conserved CD8+ T-cell epitopes. These results are discussed in the context of development of an effective anti-HIV-1 T-cell vaccine.

2. Materials and Methods

2.1. Trial HIV-CORE 004

The HIV-CORE 004 trial was conducted at the KAVI-Institute for Clinical Studies (KAVI-ICR), Kangemi site, Nairobi, Kenya between April 2014 and August 2015 and recruited healthy adults of low risk of HIV-1 infection. It had all the appropriate ethics committee and regulatory approvals and was conducted according to the principles of the Declaration of Helsinki (2008) and complied with the International Conference on Harmonization Good Clinical Practice guidelines as reported previously [60]. The trial is registered on the Pan African Clinical Trials Registry under reference PACTR201403000794397 and its ClinicalTrials.gov Identifier is NCT02099994.

2.2. Cryopreserved PBMC Samples

The PBMC used in this study were cryopreserved samples from HIV-CORE 004 and subsequently held under Human Tissue Act license 12217 at The Jenner Institute, University of Oxford, Oxford, UK. For epitope mapping, the samples used were from peak responses or the nearest available sample (between 2 and 12 weeks after vaccination).

2.3. Peptides

HIVconsv-matched 15-mer peptides overlapping by 11 amino acids (aa) >80% pure (Ana Spec, San Jose, CA, USA, 95131) and their shorter derivatives (Synpeptide, Shanghai, China) were reconstituted to 10-40 mg/mL in dimethyl sulfoxide (DMSO) (Sigma Aldridge, Pool, UK) and diluted to working stock solutions of 4 mg/mL in PBS as described previously [50].

2.4. HLA-Transfected Cell Lines

Open-reading frames coding for HLA-A*03:01, HLA-B*07:05 and HLA-B*57:03 were amplified with BamHI-Kozak-sequence and XbaI sites at their 5′- and 3′-ends, respectively, by PCR using cDNA synthesized from RNA of HLA-positive donors. The PCR products were cloned using TOPO TA or Zero Blunt Topo cloning kit (Invitrogen, Carlsbad, CA, USA). The BamHI-XbaI 1.1-kbp fragment were inserted into the same sites of the pcDNA3.1/Neo+ expression plasmid (Invitrogen, Carlsbad, CA, USA), which was used to generate 721.221-CD4 cells stably expressing HLA-A*03:01, HLA-B*07:05, or HLA-B*57:03. C1R cells expressing HLA-B*15:01 [64] and HLA-B*53:01 [65] were generated previously.

2.5. Cultured IFN-γ ELISPOT Assay

For epitope mapping, short-term cell lines (STCL) were generated from PBMC cultured in vitro for 10 days with peptide pools containing 30–33 peptides at 1.5 µg/mL per peptide in R10 culture medium (RPMI 1460 supplemented with 10% FBS, 2 mM L-glutamine, 1 mM sodium pyruvate, 10 mM HEPES, and penicillin–streptomycin antibiotics (Sigma-Aldrich, St. Louis, MO, USA)). Interleukin (IL)-7 at 25 ng/mL (Peprotech, London, UK.) was added at the start of the culture and IL-2 at 100 IU/mL was added on days 3 and 7. After 10 days, the cells were washed and rested in medium without IL-2 (R&D Systems) for at least 24 h. For the ELISPOT assay, each 15-mer peptide was tested in duplicate and the assay performed using 4.0 × 104 cells per well as described previously 3.

2.6. Intracellular Cytokine Staining (ICS) Assay and HLA Restriction

STCL were generated as above using a single “parental” 15-mer peptide. To examine optimal peptide length and their HLA restriction, 721.221 or C1R cells transfected with a single HLA class I allele were pulsed with short peptides and used as antigen-presenting cells for stimulation of STCL effectors in an ICS assay, which also confirmed that the responding T cells were CD8+ [62]. For tetramer reactivity, 106 STCL cells were stained with pre-titrated amounts of PE-conjugated tetramer (NIH Tetramer Facility, Emory University Vaccine Center, Atlanta, GA, USA) in FACS tubes at room temperature for 10 min. For both HLA and tetramer procedures, a mix of anti-CD8 FITC, anti-CD4 PE, anti-CD3 PE-CF594, anti-IFN-γ V450 (BD Biosciences, Wokingham, UK), and anti-CD3 ECD (Beckman-Coulter, High Wycombe, UK) monoclonal antibodies (mAbs), and LIVE/DEAD fixable cell stain Aqua were added and the tubes incubated for a further 20 min at room temperature. The cells were washed with FACS buffer, fixed with 1% paraformaldehyde prior to analysis, acquired by a Fortessa flow cytometer (Becton-Dickinson, Franklin lakes, NJ, USA) and analyzed using the FlowJo software (Tree Star).

2.7. In Vitro Assay for Peptide-HLA-Class I Complex Formation

The interaction of some peptides with HLA-A*02:01 was validated using an easYmer kit (Immune Aware Aps, Virum, Denmark) following the manufacturer’s instructions. This kit provided peptide-receptive and biotinylated HLA-A*02:01 heavy chains and peptide NLVPMVATV as a positive control. The assay captures refolded HLA class I complexes on streptavidin-coated microbeads (Spherotech, Saxon Europe, Kelso, UK) and quantifies them using PE-conjugated anti-β2-microglobulin (Insight Biotechnology, Wembley, UK).

2.8. Cytotoxicity Assay

A flow cytometry-based assay was used to measure the cytotoxic T lymphocyte (CTL) effector function of the YV9 clone [61]. For the CTL assay, autologous B-lymphoblastoid cell lines (B-LCL), were the targets. The Cell Trace 5 (6)-carboxyfluorescein diacetate N-succinimidyl ester (CFSE) (Molecular Probes; Invitrogen, Carlsbad, CA, USA) -labelled cells pulsed with peptide at concentrations ranging 0.01–10.0 µM or unpulsed were mixed with Far Red-labelled BCL and the CD8+ clone at 105 cells at 1:1:1 ratio in a round-bottom microtiter plate. After culture overnight, the cells were washed and stained with LIVE/DEAD Aqua (Thermo Fisher, Waltham, MA, USA) and fluorescence measured using an LSRII flow cytometer (Becton-Dickinson, Franklin lakes, NJ, USA). The ratio of CFSE: Far Red labelled cells in the absence (Eo) and presence (E+) of effector CD8+ T cells was used to measure the cytotoxicity using the equation: % Lysis = Eo − E+/Eo × 100.

3. Results

3.1. The Study Subjects and Vaccination

Healthy, HIV-1/2-negative adults in Nairobi, Kenya recruited into trial HIV-CORE 004 received a combination of HIVconsv and GRIN candidate HIV-1 vaccines as described [60]. The HIVconsv immunogen was assembled from conserved regions of HIV-1 proteins and its gene was delivered by a plasmid DNA (pSG2.HIVconsv or D) and replication-deficient poxvirus modified vaccinia virus Ankara (MVA.HIVconsv or M) [50,53]. The GRIN immunogen was a fusion of full-length Gag, Reverse Transcriptase (RT), Integrase and Nef, and was delivered by replication-deficient human adenovirus serotype 35 (Ad35-GRIN or A) [66,67]. Ad35-GRIN substituted for the simian adenovirus serotype 63-vectored HIVconsv vaccine used in previous trials [50,56,57,58,59], to which we lost freedom to operate following GlaxoSmithKline acquisition of the ChAdV63 vaccine vector. GRIN had an extensive sequence match to HIVconsv as GRIN covered almost 80% of all HIVconsv (Figure 1) [60].

3.2. HLA Allele Frequency in Nairobi Volunteers

Trial HIV-CORE 004 recruited 72 healthy, HIV-1/2-negative individuals, whose HLA class I A, B, and C alleles, the regimen they received and their total peak fresh ex vivo IFN-γ ELISPOT assay magnitude of HIVconsv-specific T cells are shown in Table S1. The HLA genotype (Figure 2, left) and HLA supertype (Figure 2, right) frequencies were determined and the most frequent alleles in this cohort for the three loci were found to be HLA-A*68:02 and A*02:01, HLA-B*15:03 and B*53:01, and HLA-C*06:02 and C*04:01, respectively, out of a total of 83 alleles (29A, 34B, and 20C).

3.3. Definition of Optimal CD8+ T-Cell Epitopes Restricted by African HLAs

The primary trial readout used 6 pools of 15-mer peptides overlapping by 11 aa (15/11) spanning the entire length of the HIVconsv immunogen in a fresh ex vivo IFN-γ ELISPOT assay. As previously reported, the AM, DDDAM and DeDeDeAM regimens induced median frequencies of 2158, 3590, and 2369 spot-forming unit (SFU)/106 PBMC, respectively, of HIVconsv-specific T cells, which recognized median of 6 out of 6 peptide pools [60]. Here, the primary aim was to define the specificity rather than magnitude of vaccine-elicited T cells for frequent African HLA alleles. PBMC were first re-stimulated in vitro with individual previously identified, strongly positive peptide pools for 10 days to generate STCL. These STCL were then tested against individual 15-mer peptides contained in the pool in an IFN-γ ELISPOT assay. Many peptides were recognized by more than one individual and the most frequently targeted peptides were contained in both the HIVconsv and GRIN immunogen sequences (Figure 3). In the second stage, STCL were expanded from PBMC using individual “parental” 15-mers and assayed against progressively shorter peptides in a polychromatic intracellular cytokine staining (ICS) assay. The HLA restriction of optimal peptides was confirmed by IFN-γ production by CD3+CD8+ cells of the STCL using peptide-pulsed B-LCL, 721.221 and/or C1R cell lines transfected with a single HLA-class I molecule (see Figure S1 for gating strategy) and reactivity with HLA-peptide tetrameric complexes. Results are presented in the order of the parental 15-mer overlapping peptides in Figures S2–S8 and summarized in Table 1. Overall, we newly defined and/or confirmed 21 optimal-length CD8+ T cell-stimulatory epitopes and 8 HLA-epitope pairs, of which 4 pairs have not been reported previously.

3.4. Peptide FF9 Stimulates Functional T-Cell Response via Three HLA Supertypes

Vaccine recipient 873 (A*03:01, A*26:01; B*15:01, B*57:03, C*04:01, C*07:01) responded to peptide HC078 YFSVPLDEGFRKYTA. Using HC078-expanded STCL, we found the strongest reactive peptide to be FSVPLDEGF (FF9), while a cluster of longer HC078 derivatives YT14, YY13, YK12, YR11, and YF10 also induced a functional T-cell response (Figure S4a). Using HLA-transfected 721.221 cells, FF9 presented by HLA-B*57:03 stimulated strongly effector cells. At the same time, HLA-A*03:01 and HLA-B*15:01-expressing cells loaded with FF9 also delivered although weaker, but a definite positive IFN-γ production signal (Figure 4a). This was further supported by peptide titration experiments (Figure 4b) and HLA/peptide tetramer reactivity for HLA-B*57:03/FF9 and HLA-B*15:01/FF9 (an HLA-A*03:01/FF9 tetramer was not available) (Figure 4c). While a potentially CD3+CD8+CD57+ NKT cell-mediated signal was detected within the HC078 STCL FF9-responsive population, a definite “classical” CD8+ T-cell component of the IFN-γ production through all restricting HLA molecules was observed (Figure S4b).

3.5. Heterogeneity of Vaccine-Elicited T-Cell Responses Restricted through HLA-A*02:01

Next, we assessed the heterogeneity of epitopes targeted in the context of a common allele HLA-A*02:01 and to what extend differences in the host genetic background impacted responses on this common allele. To this end, we tested 18 epitopes reported in the National Laboratory HIV Molecular Immunology Database (LANL-HMID) to be restricted by HLA-A2/A*02:01 solely or as well as by other HLAs [68,69] and having at least 2 out of 3 anchor residues at P2, P6 and P9 of 9-mer epitopes of the HLA-A*02:01-binding motif [21,70]. Six HLA-A*02:01-positive volunteers 807, 813, 857, 860, 884, and 889 with sufficient cryopreserved samples available were included in this analysis. Even though the optimal, truly HLA-A*02:01-presented epitope has not been defined for some of them, the six HLA-1*02:01-positive individuals recognized median (range) of 3.5 (0–6) out of these 18 tested HIVconsv epitopes (Figure 5a). Furthermore, T cells of volunteers 813, 857, and 884 also responded, albeit with variable strength, to frequently occurring natural variants of epitopes EILKDPVHGV, KAFSPEVIPMF, and FLGKIWPS, respectively (Figure 5b).

3.6. Suboptimal In Vivo Response Rescued by Index Epitope Boost

To increase the safety of the GRIN immunogen, enzymatic activity of RT was inhibited by a substitution of two bulky, charged glutamic acid residues in the RT active site MDDL by smaller, hydrophobic alanines to generate MAAL [66,67]. In contrast, HIVconsv encoded the MDDL wildtype sequence. This provided a unique opportunity to assess in vivo induction of T cells to and in vitro functional recognition of a well-defined HLA-A*02:01-restricted index epitope YQYMDDLYV (YV9) [38,43,61,62,69,71,72,73] in DNA.HIVconsv and MVA.HIVconsv, and its alanine variant delivered by Ad35-GRIN. Responses were compared after DeDeDe, DeDeDeA, and DeDeDeAM in two HLA-A*02:01-positive participants 884 and 889. Only participant 884 showed a response to this region: stimulation with the mutant HC093 (K)NPEIVIYQYMAALYV peptide expanded STCL only in samples obtained after the full DeDeDeAM regimen, but not after DeDeDeA, although DeDeDeA-primed and HC093-expanded STCL recognized YQYMAALYV. Thus, Ad35-GRIN/MAAL primed T cells required a boost with the index epitope in MVA.HIVconsv/MDDL for their in vivo expansion (Figure 6a). While both YQYMDDLYV and YQYMAALYV bound HLA-A*02:01 with same efficiency (Figure 6b), the engagement of YQYMAALYV with TCR of a YQYMDDLYV-grown T-cell clone failed completely in a killing assay (Figure 6c). Another variant CQYMDDLYV present in 1.32% of HIV sequences in the LANL-HIV Sequence Database was capable of avid binding to the HLA-A*02:01 molecule and stimulating a T-cell clone killing, albeit impaired.

4. Discussion

Definition of CTL epitopes provides important insights into the mechanisms of host responses to HIV-1 and HIV-1′s adaptation to T-cell surveillance [14,31,74,75,76,77]. It also critically informs development of an effective HIV-1 vaccine [62,63,78,79]. In the course of this work, we characterized a number of CD8+ T-cell responses specific for HIV-1 conserved protein regions induced by the HIVconsv vaccines [53] in adult HIV-1/2-negative HIV-CORE 004 participants in Nairobi [60]. First in 72 trial participants, 83 HLA alleles were detected and their frequencies and those of the corresponding HLA supertypes were established. These results concurred well with previous reports on allele frequencies in this population and were distinct from Caucasian and Japanese populations [40,41,80,81,82,83,84]. It was suggested that while frequencies of HLA alleles might vary dramatically among different ethnicities and global locations, the frequency of each HLA allele supertype is remarkably constant [20,21]. Overall, we confirmed many already reported optimal CD8+ T-cell epitopes, for some of them identified their alternative HLA-restriction, and defined 4 novel HLA-peptide pairs, namely HLA-B*53:01/SPIETVPVK, HLA-B*07:05/SPAIFQSSMTK, HLA-B*53:01/SPAIFQSSM, and HLA-B*07:05/SPAIFQSSM [85]. These pairs represent additions to the cumulative knowledge of well-defined epitopes in this geographic region.
Volunteer 873 presented epitope FSVPLDEGF (FF9) by 3 different HLA molecules HLA-A*03:01, HLA-B*15:01 and HLA-B*57:03. These alleles belong to different HLA supertypes of A03, B62, and B58, respectively. A closer inspection of the main anchor residues in position P2 and P9 confirmed expected aa fitting into MHC pockets B and F for the strongest HLA-B*57:03 [86], but only one and none for the weaker binding alleles making the involvement of secondary anchors likely. In line with these observations, tetramer reactivity yielded different frequencies of FF9 specific cells for responses restricted by HLA-B*57:03 and HLA-B*15:01 (Figure 4c). Typical interpretation of these data would be that FF9 is presented by HLA-B*57:03, whereas the biological significance of weak HLA/peptide interactions remains unclear. However, sometimes even very weak protein–protein interactions may have biological consequences, such as those implicated in experimental challenge/protection by HLA-E-restricted CD8+ T cells [87,88,89,90,91] and CD4+ T-cell function [92,93]. It is of note that MHC-peptide affinity alone does not correlate with frequency of peptide recognition [94] nor necessarily with magnitude of responses [46] or functionality [18]. In addition, for stimulation through HLA-A*03:01, HLA-B*15:01 and HLA-B*57:03, 33.5%, 37.2%, and 25.8% subsets of the IFN-γ-producing cells in the HC078 STCL were CD3+CD8+CD57+ NKT cells (Figure S4b). The potential contribution of such a population to virus control will need to be further explored.
Six volunteers recognized median of 3.5 out of 18 tested epitopes recorded in the LANL-HMID to be presented through HLA-A*02:01/A2 (Figure 5a). Of note, the precision of the general listings in the LANL-HMID table of CTL/CD8+ Epitope Summary [85] might sometimes be questionable, which has led to the creation and regular updates of a more stringent “A” list of the best defined epitopes [85]. On the other hand, highly personalized epitope hierarchies, i.e., absence of responses to some, is not unexpected given the complexity of the antigen processing and presentation pathways [95,96,97] competition from different sets of other HLA molecules in each individual, different T-cell repertoires and, perhaps, even differences in co-infections/microbiome [98,99,100,101]. We also demonstrated that a T-cell response to an epitope with suboptimal affinity for the TCR(s) could be rescued by a vaccination with its index epitope (Figure 6). Several mechanisms could have contributed to this observation including a limited room for TCR cross-reactivity even for the perceived stringent HLA class I-peptide recognition [102] combined with multiple T-cell clones recruited into the response to this epitope [103], some of which were of too low frequency to be detected after the Ad35-GRIN prime alone. This may be another indication that even weak protein–protein interactions influence the final outcome and supports vaccines employing multivalent immunogens [69,104,105,106], although further investigations are required to generalize this observation.

5. Conclusions

In conclusion, we believe our study is a small, but significant contribution to the knowledge and understanding of likely protective CD8+ T-cell responses to HIV-1 infection in sub-Saharan Africa, where consequences of the HIV-1 epidemic still represent one of the major health and social challenges, and where an effective vaccine is urgently needed.

Supplementary Materials

The following are available online at https://www.mdpi.com/2076-393X/8/2/260/s1, Figure S1: Gating strategy for determination of HLA class I-restriction following peptide-specific restimulation, Figure S2: Responses to HC014 EWDR-IYKRWIILGLN (Gag 261–271), Figure S3: Responses to HC049 (K)NFPISPIETVPVKLK / HC050 SPIETVPVKLKPGMD (Pol 158–171), Figure S4: Responses to HC078 YFSVPLDEGFRKYTA (Pol 270–284, Figure S5. Responses to HC088 GSPAIFQSSMTKILE (Pol 311–325), Figure S6. Responses to HC164 VQMAVFIHNFKRKGGI (Pol 891–905), Figure S7: Responses to HC176 VVPRRKAKIIRDYGK (Pol 974–988), Figure S8: Responses to HC177 RKAKIIRDYGKQMAG (Pol 978–992), Table S1: HIV-1/2-negative vaccine recipients in HIV-CORE 004 used for this study, their tissue types, regimens they received, and their peak IFN-γ ELISPOT assay responses to the conserved regions of the HIVconsv immunogen induced by vaccination.

Author Contributions

Conceptualization, T.H.; methodology, Y.S.M., N.J.B., F.S., Z.H., N.M., H.M., P.H., T.A., C.B., M.T. and T.H.; software, Y.S.M., N.J.B. and T.H.; formal analysis, Y.S.M., N.J.B., T.H., N.M., S.S.-A., A.L., M.T., and C.B.; investigation, Y.S.M., N.J.B., N.M., H.M., P.E.F., W.J., and G.M.; resources, T.H., and M.T.; data curation, Y.S.M., N.J.B., N.M., T.H.; writing—original draft preparation, N.J.B. and T.H.; writing—review and editing, all authors.; supervision, T.H., and M.T.; project administration, A.C., and T.H.; funding acquisition, T.H., M.T., and P.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was jointly funded by the Medical Research Council (MRC) UK and the UK Department for International Development (DFID) under the MRC/DFID Concordat agreements (G1001757 and MR/N023668/1) to T.H.; the European and Developing Countries Clinical Trials Partnership (EDCTP) Strategic Primer Grant no. SP.2011.41304.022 to T.H. and W.J.; grants-in-aid (15fk0410019h0001, 16fk0410202h0002, and17fk0410302h0003) for AIDS Research from AMED and grants-in-aid (26293240, 17K10021) for scientific research from the Ministry of Education, Science, Sports, and Culture, Japan to M.T.; and the International AIDS Vaccine Initiative (IAVI) and through IAVI by the support of the United States Agency for International Development (USAID) and other donors. The full list of IAVI donors is available at http://www.iavi.org. The contents of this manuscript are the responsibility of IAVI and do not necessarily reflect the views of USAID or the US Government. Some sub-analyses were partly funded by the European Union’s Horizon 2020 Research and Innovation programme under grant agreement no. 681137-EAVI2020 awarded to T.H. and C.B. T.H. is a Jenner Institute Investigator.

Acknowledgments

The following tetramer reagents were obtained through the NIH Tetramer Core Facility: HLA-A*03:01 HIV POL 313–321 AIFQSSMTK PE-Label; HLA-B*07:05 HIV POL 311–321 SPAIFQSSMTK PE-Label; HLA-B*15:01 HIV POL 271-279 FSVPLDEGF PE-Label; HLA-B*53:01 HIV POL 311–319 SPAIFQSSM PE-Label; and HLA-B*57:03 HIV POL 271–279 FSVPLDEGF PE-Label. The authors would like to thank the trial HIV-CORE 004 volunteers for making this study possible and acknowledge the contributions of the HIV-CORE 004 study team: Jan De Bont, Nicodemus Kisengese, Anne Gumbe, Paramesh Chetty, Kristen Syversten, Leslie Nielsen, Carl Verlinde, Jennifer Lehrman, Megan McBride, Zablon Omungo, Eddy Sayeed, Jim Ackland and Hansi Dean.

Conflicts of Interest

The authors declare no conflict of interest with the subject matter other than T Hanke is a coinventor on patents WO06123156, WO98/056919, PCT/US2014/058422, and EP14846993.5. The authors have no other relevant affiliations or financial involvement with any organization or entity with a financial interest in or financial conflict with the subject matter or materials discussed in the manuscript apart from those disclosed. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Townsend, A.; Bodmer, H. Antigen recognition by class I-restricted T lymphocytes. Annu. Rev. Immunol. 1989, 7, 601–624. [Google Scholar] [CrossRef] [PubMed]
  2. Trowsdale, J.; Campbell, R.D. Complexity in the major histocompatibility complex. Eur. J. Immunogenet. 1992, 19, 45–55. [Google Scholar] [CrossRef] [PubMed]
  3. Bjorkman, P.; Parham, J.P. Structure, function, and diversity of class I major histocompatibility complex molecules. Annu. Rev. Biochem. 1990, 59, 253–288. [Google Scholar] [CrossRef] [PubMed]
  4. Hewitt, E.W. The MHC class I antigen presentation pathway: Strategies for viral immune evasion. Immunology 2003, 110, 163–169. [Google Scholar] [CrossRef] [PubMed]
  5. Fellay, J.; Frahm, N.; Shianna, K.V.; Cirulli, E.T.; Casimiro, D.R.; Robertson, M.N.; Haynes, B.F.; Geraghty, D.E.; McElrath, M.J.; Goldstein, D.B. Host genetic determinants of T cell responses to the MRKAd5 HIV-1 gag/pol/nef vaccine in the step trial. J. Infect. Dis. 2012, 203, 773–779. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Fellay, J.; Shianna, K.V.; Ge, D.; Colombo, S.; Ledergerber, B.; Weale, M.; Zhang, K.; Gumbs, C.; Castagna, A.; Cossarizza, A.; et al. A whole-genome association study of major determinants for host control of HIV-1. Science 2007, 317, 944–947. [Google Scholar] [CrossRef] [Green Version]
  7. Carrington, M.; O’Brien, S.J. The influence of HLA genotype on AIDS. Annu. Rev. Med. 2003, 54, 535–551. [Google Scholar] [CrossRef] [Green Version]
  8. Dalmasso, C.; Carpentier, W.; Meyer, L.; Rouzioux, C.; Goujard, C.; Chaix, M.L.; Lambotte, O.; Avettand-Fenoel, V.; Le Clerc, S.; de Senneville, L.D.; et al. Distinct genetic loci control plasma HIV-RNA and cellular HIV-DNA levels in HIV-1 infection: The ANRS Genome Wide Association 01 study. PLoS ONE 2008, 3, e3907. [Google Scholar] [CrossRef] [Green Version]
  9. Goonetilleke, N.; Liu, M.K.; Salazar-Gonzalez, J.F.; Ferrari, G.; Giorgi, E.; Ganusov, V.V.; Keele, B.F.; Learn, G.H.; Turnbull, E.L.; Salazar, M.G.; et al. The first T cell response to transmitted/founder virus contributes to the control of acute viremia in HIV-1 infection. J. Exp. Med. 2009, 206, 1253–1272. [Google Scholar] [CrossRef]
  10. Salazar-Gonzalez, J.F.; Salazar, M.G.; Keele, B.F.; Learn, G.H.; Giorgi, E.E.; Li, H.; Decker, J.M.; Wang, S.; Baalwa, J.; Kraus, M.H.; et al. Genetic identity, biological phenotype, and evolutionary pathways of transmitted/founder viruses in acute and early HIV-1 infection. J. Exp. Med. 2009, 206, 1273–1289. [Google Scholar] [CrossRef]
  11. Juarez-Molina, C.I.; Payne, R.; Soto-Nava, M.; Avila-Rios, S.; Valenzuela-Ponce, H.; Adland, E.; Leitman, E.; Brener, J.; Muenchhoff, M.; Branch, S.; et al. Impact of HLA selection pressure on HIV fitness at a population level in Mexico and Barbados. J. Virol. 2014, 88, 10392–10398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Kawashima, Y.; Pfafferott, K.; Frater, J.; Matthews, P.; Payne, R.; Addo, M.; Gatanaga, H.; Fujiwara, M.; Hachiya, A.; Koizumi, H.; et al. Adaptation of HIV-1 to human leukocyte antigen class I. Nature 2009, 458, 641–645. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Martinez-Picado, J.; Prado, J.G.; Fry, E.E.; Pfafferott, K.; Leslie, A.; Chetty, S.; Thobakgale, C.; Honeyborne, I.; Crawford, H.; Matthews, P.; et al. Fitness cost of escape mutations in p24 Gag in association with control of human immunodeficiency virus type 1. J. Virol. 2006, 80, 3617–3623. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Payne, R.; Muenchhoff, M.; Mann, J.; Roberts, H.E.; Matthews, P.; Adland, E.; Hempenstall, A.; Huang, K.H.; Brockman, M.; Brumme, Z.; et al. Impact of HLA-driven HIV adaptation on virulence in populations of high HIV seroprevalence. Proc. Nat. Acad. Sci. USA 2014, 111, E5393–E5400. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Yue, L.; Pfafferott, K.J.; Baalwa, J.; Conrod, K.; Dong, C.C.; Chui, C.; Rong, R.; Claiborne, D.T.; Prince, J.L.; Tang, J.; et al. Transmitted virus fitness and host T cell responses collectively define divergent infection outcomes in two HIV-1 recipients. PLoS Pathog. 2015, 11, e1004565. [Google Scholar] [CrossRef] [Green Version]
  16. Haynes, B.F.; Pantaleo, G.; Fauci, A.S. Toward an understanding of the correlates of protective immunity to HIV infection. Science 1996, 271, 324–328. [Google Scholar] [CrossRef]
  17. Kaur, G.; Mehra, N. Genetic determinants of HIV-1 infection and progression to AIDS: Immune response genes. Tissue Antigens 2009, 74, 373–385. [Google Scholar] [CrossRef]
  18. Leslie, A.; Price, D.A.; Mkhize, P.; Bishop, K.; Rathod, A.; Day, C.; Crawford, H.; Honeyborne, I.; Asher, T.E.; Luzzi, G.; et al. Differential selection pressure exerted on HIV by CTL targeting identical epitopes but restricted by distinct HLA alleles from the same HLA supertype. J. Immunol. 2006, 177, 4699–4708. [Google Scholar] [CrossRef] [Green Version]
  19. Leslie, A.J.; Pfafferott, K.J.; Chett, P.; Draenert, R.; Addo, M.M.; Feeney, M.; Tang, Y.; Holmes, E.C.; Allen, T.; Prado, J.G.; et al. HIV evolution: CTL escape mutation and reversion after transmission. Nat. Med. 2004, 10, 282–289. [Google Scholar] [CrossRef]
  20. Sette, A.; Sidney, J. Nine major HLA class I supertypes account for the vast preponderance of HLA-A and -B polymorphism. Immunogenetics 1999, 50, 201–212. [Google Scholar] [CrossRef]
  21. Sidney, J.; Peters, B.; Frahm, N.; Brander, C.; Sette, A. HLA class I supertypes: A revised and updated classification. BMC Immunol. 2008, 9, 1. [Google Scholar] [CrossRef] [Green Version]
  22. Trachtenberg, E.; Korber, B.; Sollars, C.; Kepler, T.B.; Hraber, P.T.; Hayes, E.; Funkhouser, R.; Fugate, M.; Theiler, J.; Hsu, Y.S.; et al. Advantage of rare HLA supertype in HIV disease progression. Nat. Med. 2003, 9, 928–935. [Google Scholar] [CrossRef]
  23. Bodmer, W.F. Evolutionary significance of the HL-A system. Nature 1972, 237, 139–145. [Google Scholar] [CrossRef] [PubMed]
  24. Hill, A.V.; Allsopp, C.E.; Kwiatkowski, D.; Anstey, N.M.; Twumasi, P.; Rowe, P.A.; Bennett, S.; Brewster, D.; McMichael, A.J.; Greenwood, B.M. Common west African HLA antigens are associated with protection from severe malaria. Nature 1991, 352, 595–600. [Google Scholar] [PubMed]
  25. Hughes, A.L.; Ota, T.; Nei, M. Positive Darwinian selection promotes charge profile diversity in the antigen-binding cleft of class I major-histocompatibility-complex molecules. Mol. Biol. Evol. 1990, 7, 515–524. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Little, A.M.; Parham, P. Polymorphism and evolution of HLA class I and II genes and molecules. Rev. Immunogenet. 1999, 1, 105–123. [Google Scholar]
  27. Demers, K.R.; Reuter, M.A.; Betts, M.R. CD8(+) T-cell effector function and transcriptional regulation during HIV pathogenesis. Immunol. Rev. 2013, 254, 190–206. [Google Scholar] [CrossRef] [Green Version]
  28. Eller, M.A.; Goonetilleke, N.; Tassaneetrithep, B.; Eller, L.A.; Costanzo, M.C.; Johnson, S.; Betts, M.R.; Krebs, S.J.; Slike, B.M.; Nitayaphan, S.; et al. Expansion of Inefficient HIV-Specific CD8 T Cells during Acute Infection. J. Virol. 2016, 90, 4005–4016. [Google Scholar] [CrossRef] [Green Version]
  29. Makedonas, G.; Betts, M.R. Living in a house of cards: Re-evaluating CD8+ T-cell immune correlates against HIV. Immunol. Rev. 2011, 239, 109–124. [Google Scholar] [CrossRef]
  30. Walker, B.; McMichael, A. The T-Cell Response to HIV. Cold Spring Harb. Perspect. Med. 2012, 2, 1–19. [Google Scholar] [CrossRef] [Green Version]
  31. Walker, B.D.; Korber, B.T. Immune control of HIV: The obstacles of HLA and viral diversity. Nat. Immunol. 2001, 2, 473–475. [Google Scholar] [CrossRef] [PubMed]
  32. Betts, M.R.; Ambrozak, D.R.; Douek, D.C.; Bonhoeffer, S.; Brenchley, J.M.; Casazza, J.P.; Koup, R.A.; Picker, L.J. Analysis of total human immunodeficiency virus (HIV)-specific CD4(+) and CD8(+) T-cell responses: Relationship to viral load in untreated HIV infection. J. Virol. 2001, 75, 11983–11991. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Betts, M.R.; Nason, M.C.; West, S.M.; De Rosa, S.C.; Migueles, S.A.; Abraham, J.; Lederman, M.M.; Benito, J.M.; Goepfert, P.A.; Connors, M.; et al. HIV nonprogressors preferentially maintain highly functional HIV-specific CD8+ T cells. Blood 2006, 107, 4781–4789. [Google Scholar]
  34. Ferrari, G.; Korber, B.; Goonetilleke, N.; Liu, M.K.; Turnbull, E.L.; Salazar-Gonzalez, J.F.; Hawkins, N.; Self, S.; Watson, S.; Betts, M.R.; et al. Relationship between functional profile of HIV-1 specific CD8 T cells and epitope variability with the selection of escape mutants in acute HIV-1 infection. PLoS Pathog 2011, 7, e1001273. [Google Scholar] [CrossRef] [PubMed]
  35. Nguyen, S.; Deleage, C.; Darko, S.; Ransier, A.; Truong, D.P.; Agarwal, D.; Japp, A.S.; Wu, V.H.; Kuri-Cervantes, L.; Abdel-Mohsen, M.; et al. Elite control of HIV is associated with distinct functional and transcriptional signatures in lymphoid tissue CD8(+) T cells. Sci. Transl. Med. 2019, 11. [Google Scholar] [CrossRef] [PubMed]
  36. Price, D.A.; Brenchley, J.M.; Ruff, L.E.; Betts, M.R.; Hill, B.J.; Roederer, M.; Koup, R.A.; Migueles, R.A.; Gostick, E.; Wooldridge, L.; et al. Avidity for antigen shapes clonal dominance in CD8+ T cell populations specific for persistent DNA viruses. J. Exp. Med. 2005, 202, 1349–1361. [Google Scholar] [CrossRef] [PubMed]
  37. Frahm, N.; Kiepiela, P.; Adams, S.; Linde, C.H.; Hewitt, H.S.; Sango, K.; Feeney, M.E.; Addo, M.M.; Lichterfeld, M.; Lahaie, M.P.; et al. Control of human immunodeficiency virus replication by cytotoxic T lymphocytes targeting subdominant epitopes. Nat. Immunol. 2006, 7, 173–178. [Google Scholar] [CrossRef]
  38. Kunwar, P.; Hawkins, N.; Dinges, W.L.; Liu, Y.; Gabriel, E.E.; Swan, D.A.; Stevens, C.E.; Maenza, J.; Collier, A.C.; Mullins, J.I.; et al. Superior Control of HIV-1 Replication by CD8+ T Cells Targeting Conserved Epitopes: Implications for HIV Vaccine Design. PLoS ONE 2013, 8, e64405. [Google Scholar] [CrossRef] [Green Version]
  39. Mothe, B.; Llano, A.; Ibarrondo, J.; Daniels, M.; Miranda, C.; Zamarreno, J.; Bach, V.; Zuniga, R.; Perez-Alvarez, S.; Berger, C.T.; et al. Definition of the viral targets of protective HIV-1-specific T cell responses. J. Transl. Med. 2011, 9, 208. [Google Scholar] [CrossRef] [Green Version]
  40. Murakoshi, H.; Akahoshi, T.; Koyanagi, M.; Chikata, T.; Naruto, T.; Maruyama, R.; Tamura, Y.; Ishizuka, N.; Gatanaga, H.; Oka, S.; et al. Clinical Control of HIV-1 by Cytotoxic T Cells Specific for Multiple Conserved Epitopes. J. Virol. 2015, 89, 5330–5339. [Google Scholar] [CrossRef] [Green Version]
  41. Murakoshi, H.; Zou, C.; Kuse, N.; Akahoshi, T.; Chikata, T.; Gatanaga, H.; Oka, S.; Hanke, T.; Takiguchi, M. CD8+ T cells specific for conserved, cross-reactive Gag epitopes with strong ability to suppress HIV-1 replication. Retrovirology 2018, 15, 46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Ondondo, B.; Murakoshi, H.; Clutton, G.; Abdul-Jawad, S.; Wee, E.G.; Gatanaga, H.; Oka, S.; McMichael, A.J.; Takiguchi, M.; Korber, B.; et al. Novel Conserved-region T-cell Mosaic Vaccine With High Global HIV-1 Coverage Is Recognized by Protective Responses in Untreated Infection. Mol. Ther. 2016, 24, 832–842. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Zou, C.; Murakoshi, H.; Kuse, N.; Akahoshi, T.; Chikata, T.; Gatanaga, H.; Oka, S.; Hanke, T.; Takiguchi, M. Effective Suppression of HIV-1 Replication by Cytotoxic T Lymphocytes Specific for Pol Epitopes in Conserved Mosaic Vaccine Immunogens. J. Virol. 2019, 93, e02142–e02218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Altfeld, M.; Addo, M.M.; Rosenberg, E.S.; Hecht, F.M.; Lee, P.K.; Vogel, M.; Yu, X.G.; Draenert, R.; Johnston, M.N.; Strick, D.; et al. Influence of HLA-B57 on clinical presentation and viral control during acute HIV-1 infection. AIDS 2003, 17, 2581–2591. [Google Scholar]
  45. Altfeld, M.; Kalife, E.T.; Qi, Y.; Streeck, H.; Lichterfeld, M.; Johnston, M.N.; Burgett, N.; Swartz, M.E.; Yang, A.; Alter, G.; et al. HLA Alleles Associated with Delayed Progression to AIDS Contribute Strongly to the Initial CD8(+) T Cell Response against HIV-1. PLoS. Med. 2006, 3, e403. [Google Scholar] [CrossRef]
  46. Bihl, F.; Frahm, N.; Di Giammarino, L.; Sidney, J.; John, M.; Yusim, K.; Woodberry, T.; Sango, K.; Hewitt, H.S.; Henry, L.; et al. Impact of HLA-B alleles, epitope binding affinity, functional avidity, and viral coinfection on the immunodominance of virus-specific CTL responses. J. Immunol. 2006, 176, 4094–4101. [Google Scholar]
  47. Crawford, H.; Lumm, W.; Leslie, A.; Schaefer, M.; Boeras, D.; Prado, J.G.; Tang, J.; Farmer, P.; Ndung’u, T.; Lakhi, S.; et al. Evolution of HLA-B*5703 HIV-1 escape mutations in HLA-B*5703-positive individuals and their transmission recipients. J. Exp. Med. 2009, 206, 909–921. [Google Scholar] [CrossRef]
  48. Flores-Villanueva, P.O.; Yunis, E.J.; Delgado, J.C.; Vittinghoff, E.; Buchbinder, S.; Leung, J.Y.; Uglialoro, A.M.; Clavijo, O.P.; Rosenberg, E.S.; Kalams, S.A.; et al. Control of HIV-1 viremia and protection from AIDS are associated with HLA-Bw4 homozygosity. Proc. Nat. Acad. Sci. USA 2001, 98, 5140–5145. [Google Scholar] [CrossRef] [Green Version]
  49. Goulder, P.J.; Walker, B.D. HIV and HLA class I: An evolving relationship. Immunity 2012, 37, 426–440. [Google Scholar] [CrossRef] [Green Version]
  50. Borthwick, N.; Ahmed, T.; Ondondo, B.; Hayes, P.; Rose, A.; Ebrahimsa, U.; Hayton, E.J.; Black, A.; Bridgeman, A.; Rosario, M.; et al. Vaccine-elicited human T cells recognizing conserved protein regions inhibit HIV-1. Mol. Ther. 2014, 22, 464–475. [Google Scholar] [CrossRef] [Green Version]
  51. Hanke, T. Conserved immunogens in prime-boost strategies for the next-generation HIV-1 vaccines. Expert. Opin. Biol. Ther. 2014, 14, 601–616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Hanke, T. Aiming for protective T-cell responses: A focus on the first generation conserved-region HIVconsv vaccines in preventive and therapeutic clinical trials. Expert. Rev. Vaccines 2019, 18, 1029–1041. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Letourneau, S.; Im, E.-J.; Mashishi, T.; Brereton, C.; Bridgeman, A.; Yang, H.; Dorrell, L.; Dong, T.; Korber, B.; McMichael, A.J.; et al. Design and pre-clinical evaluation of a universal HIV-1 vaccine. PLoS ONE 2007, 2, e984. [Google Scholar] [CrossRef] [PubMed]
  54. Murakoshi, H.; Kuse, N.; Akahoshi, T.; Zhang, Y.; Chikata, T.; Borghan, M.A.; Gatanaga, H.; Oka, S.; Sakai, K.; Takiguchi, M. Broad Recognition of Circulating HIV-1 by HIV-1-Specific Cytotoxic T-Lymphocytes with Strong Ability to Suppress HIV-1 Replication. J. Virol. 2019, 93, e01480–e01518. [Google Scholar] [CrossRef] [Green Version]
  55. Rolland, M.; Nickle, D.C.; Mullins, J.I. HIV-1 group M conserved elements vaccine. PLoS Pathog. 2007, 3, e157. [Google Scholar] [CrossRef] [Green Version]
  56. Borthwick, N.J.; Lane, T.; Moyo, N.; Crook, A.; Shim, J.M.; Baines, I.; Wee, E.G.; Hawkins, P.H.; Gillmore, J.D.; Hanke, T.; et al. Randomized phase I trial HIV-CORE 003: Depletion of serum amyloid P component and immunogenicity of DNA vaccination against HIV-1. PLoS ONE 2018, 13, e0197299. [Google Scholar] [CrossRef]
  57. Fidler, S.; Stohr, W.; Pace, M.; Dorrell, L.; Lever, A.; Pett, S.; Kinloch-de Loes, S.; Fox, J.; Clarke, A.; Nelson, M.; et al. Antiretroviral therapy alone versus antiretroviral therapy with a kick and kill approach, on measures of the HIV reservoir in participants with recent HIV infection (the RIVER trial): A phase 2, randomised trial. Lancet 2020, 395, 888–898. [Google Scholar] [CrossRef]
  58. Mothe, B.; Manzardo, C.; Snachez-Bernabeau, A.; Coll, P.; Moron-Lopez, S.; Puertas, M.C.; Rosas, M.; Cobarsi, P.; Escrig, N.; Perez-Alvarez, N.; et al. Therapeutic vaccination refocused T-cell responses to conserved regions of HIV-1 in early reated individuals (BCN 01 study). Lancet eClinMed 2019, 1, 65–80. [Google Scholar] [CrossRef] [Green Version]
  59. Mothe, B.; Rosas-Umbert, M.; Coll, P.; Manzardo, C.; Puertas, M.C.; Morón-López, S.; Llano, A.; Miranda, C.; Cedeño, S.; López, M.; et al. HIVconsv vaccines and romidepsin in early-treated HIV-1-infected individuals: Safety, immunogenicity and effect on the viral reservoir (study BCN 02). Front. Immunol. 2020. [Google Scholar] [CrossRef]
  60. Mutua, G.; Farah, B.; Langat, R.; Indangasi, J.; Ogola, S.; Onsembe, B.; Kopycinski, J.T.; Hayes, P.; Ashraf, A.; Borthwick, N.J.; et al. Broad HIV-1 inhibition in vitro by vaccine-elicited CD8+ T cells in African adults. Mol. Ther. Methods Clin. Dev. 2016, 3, 16061. [Google Scholar] [CrossRef]
  61. Ahmed, T.; Borthwick, N.J.; Gilmour, J.; Hayes, P.; Dorrell, L.; Hanke, T. Control of HIV-1 replication in vitro by vaccine-induced human CD8 T cells through conserved subdominant Pol epitopes. Vaccine 2016, 34, 1215–1224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Borthwick, N.; Lin, Z.; Akahoshi, T.; Llano, A.; Silva-Arrieta, S.; Ahmed, T.; Dorrell, L.; Brander, C.; Murakoshi, H.; Takiguchi, M.; et al. Novel, in-natural-infection subdominant HIV-1 CD8+ T-cell epitopes revealed in human recipients of conserved-region T-cell vaccines. PLoS ONE 2017, 12, e0176418. [Google Scholar] [CrossRef] [PubMed]
  63. Borthwick, N.; Silva-Arrieta, S.; Llano, A.; Takiguchi, M.; Brander, C.; Hanke, T. Novel Nested Peptide Epitopes Recognized by CD4(+) T Cells Induced by HIV-1 Conserved-Region Vaccines. Vaccines (Basel) 2020, 8, 28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Falk, K.; Rotzschke, O.; Takiguchi, M.; Gnau, V.; Stevanovic, S.; Jung, G.; Rammensee, H.G. Peptide motifs of HLA-B58, B60, B61, and B62 molecules. Immunogenetics 1995, 41, 165–168. [Google Scholar] [CrossRef] [PubMed]
  65. Yamamoto, J.; Kariyone, A.; Akiyama, N.; Kano, K.; Takiguchi, M. Presentation of human minor histocompatibility antigens by HLA-B35 and HLA-B38 molecules. Proc. Natl. Acad. Sci. USA 1990, 87, 2583–2587. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Mpendo, J.; Mutua, G.; Nyombayire, J.; Ingabire, R.; Nanvubya, A.; Anzala, O.; Karita, E.; Hayes, P.; Kopycinski, J.; Dally, L.; et al. A Phase I Double Blind, Placebo-Controlled, Randomized Study of the Safety and Immunogenicity of Electroporated HIV DNA with or without Interleukin 12 in Prime-Boost Combinations with an Ad35 HIV Vaccine in Healthy HIV-Seronegative African Adults. PLoS ONE 2015, 10, e0134287. [Google Scholar] [CrossRef] [Green Version]
  67. Omosa-Manyonyi, G.; Mpendo, J.; Ruzagira, E.; Kilembe, W.; Chomba, E.; Roman, F.; Bourguignon, P.; Koutsoukos, M.; Collard, A.; Voss, G.; et al. A Phase I Double Blind, Placebo-Controlled, Randomized Study of the Safety and Immunogenicity of an Adjuvanted HIV-1 Gag-Pol-Nef Fusion Protein and Adenovirus 35 Gag-RT-Int-Nef Vaccine in Healthy HIV-Uninfected African Adults. PLoS ONE 2015, 10, e0125954. [Google Scholar] [CrossRef]
  68. LANL HMID CTL/CD8+ Epitope Summary. 2020. Available online: https://www.hiv.lanl.gov/content/immunology/tables/ctl_summary.html (accessed on 3 May 2020).
  69. Abdul-Jawad, S.; Ondondo, B.; van Hateren, A.; Gardner, A.; Elliott, T.; Korber, B.; Hanke, T. Increased Valency of Conserved-mosaic Vaccines Enhances the Breadth and Depth of Epitope Recognition. Mol. Ther. 2016, 24, 375–384. [Google Scholar] [CrossRef] [Green Version]
  70. Sliz, P.; Michielin, O.; Cerottini, J.C.; Luescher, I.; Romero, P.; Karplus, M.; Wiley, D.C. Crystal structures of two closely related but antigenically distinct HLA-A2/melanocyte-melanoma tumor-antigen peptide complexes. J. Immunol. 2001, 167, 3276–3284. [Google Scholar] [CrossRef] [Green Version]
  71. Roider, J.; Meissner, T.; Kraut, F.; Vollbrecht, T.; Stirner, R.; Bogner, J.R.; Draenert, R. Comparison of experimental fine-mapping to in silico prediction results of HIV-1 epitopes reveals ongoing need for mapping experiments. Immunology 2014, 143, 193–201. [Google Scholar] [CrossRef]
  72. Samri, A.; Haas, G.; Duntze, J.; Bouley, J.M.; Calvez, V.; Katlama, C.; Autran, B. Immunogenicity of mutations induced by nucleoside reverse transcriptase inhibitors for human immunodeficiency virus type 1-specific cytotoxic T cells. J. Virol. 2000, 74, 9306–9312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Vollbrecht, T.; Eberle, J.; Roider, J.; Bühler, S.; Stirner, R.; Henrich, N.; Seybold, U.; Bogner, J.R.; Draenert, R. Control of M184V HIV-1 mutants by CD8 T-cell responses. Med. Microbiol. Immunol. 2012, 201, 201–211. [Google Scholar] [CrossRef] [PubMed]
  74. Johnson, R.P.; Trocha, A.; Buchanan, T.M.; Walker, B.D. Identification of overlapping HLA class I-restricted cytotoxic T cell epitopes in a conserved region of the human immunodeficiency virus type 1 envelope glycoprotein: Definition of minimum epitopes and analysis of the effects of sequence variation. J. Exp. Med. 1992, 175, 961–971. [Google Scholar]
  75. Johnson, R.P.; Trocha, A.; Buchanan, T.M.; Walker, B.D. Recognition of a highly conserved region of human immunodeficiency virus type 1 gp120 by an HLA-Cw4-restricted cytotoxic T-lymphocyte clone. J. Virol. 1993, 67, 438–445. [Google Scholar] [PubMed]
  76. Kloverpris, H.N.; Stryhn, A.; Harndahl, M.; Payne, R.; Towers, G.J.; Chen, F.; Riddell, L.; Walker, B.D.; Ndung’u, T.; Leslie, A.; et al. HLA-specific intracellular epitope processing shapes an immunodominance pattern for HLA-B*57 that is distinct from HLA-B*58:01. J. Virol. 2013, 87, 10889–10894. [Google Scholar] [CrossRef] [Green Version]
  77. Kloverpris, H.N.; Stryhn, A.; Harndahl, M.; van der Stok, M.; Payne, R.P.; Matthews, P.C.; Chen, F.; Riddell, L.; Walker, B.D.; Ndung’u, T.; et al. HLA-B*57 Micropolymorphism shapes HLA allele-specific epitope immunogenicity, selection pressure, and HIV immune control. J. Virol. 2012, 86, 919–929. [Google Scholar] [CrossRef] [Green Version]
  78. Llano, A.; Cedeño, S.; Arrieta, S.S.; Brander, C. Optimal HIV CTL epitopes update: Growing diversity in epitope length and HLA restriction; HIV Immunology and HIV/SIV Vaccine Databases; Theoretical Biology and Biophysics Group, Los Alamos National Laboratory: Los Alamos, NM, USA, 2019.
  79. Lano, A.; Williams, A.; Olvera, A.; Silva-Arrieta, S.; Brander, D. Best-Characterized HIV-1 CTL Epitopes: The 2013 Update; HIV Immunology and HIV/SIV Vaccine Databases; Theoretical Biology and Biophysics Group, Los Alamos National Laboratory: Los Alamos, NM, USA, 2013.
  80. Abidi, S.H.; Shahid, A.; Lakhani, L.S.; Shah, R.; Okinda, N.; Ojwang, P.; Abbas, F.; Rowland-Jones, S.; Ali, S. HIV-1 progression links with viral genetic variability and subtype, and patient’s HLA type: Analysis of a Nairobi-Kenyan cohort. Med. Microbiol. Immunol. 2014, 203, 57–63. [Google Scholar] [CrossRef]
  81. Bird, T.G.; Kaul, R.; Rostron, T.; Kimani, J.; Embree, J.; Dunn, P.P.; Bwayo, J.J.; Plummer, F.A.; Rowland-Jones, S.L.; Dong, T.; et al. HLA typing in a Kenyan cohort identifies novel class I alleles that restrict cytotoxic T-cell responses to local HIV-1 clades. AIDS 2002, 16, 1899–1904. [Google Scholar] [CrossRef] [Green Version]
  82. Cao, K.; Moormann, A.M.; Lyke, K.E.; Masaberg, C.; Sumba, O.P.; Doumbo, O.K.; Koech, D.; Lancaster, A.; Nelson, M.; Meyer, D.; et al. Differentiation between African populations is evidenced by the diversity of alleles and haplotypes of HLA class I loci. Tissue Antigens 2004, 63, 293–325. [Google Scholar] [CrossRef]
  83. Khoja, S.; Ojwang, P.; Khan, S.; Okinda, N.; Harania, R.; Ali, S. Genetic analysis of HIV-1 subtypes in Nairobi, Kenya. PLoS ONE 2008, 3, e3191. [Google Scholar] [CrossRef] [Green Version]
  84. Tang, J.; Tang, S.; Lobashevsky, E.; Myracle, A.D.; Fideli, U.; Aldrovandi, G.; Allen, S.; Musonda, R.; Kaslow, R.A.; Zambia-UAB HIV Research Project. Favorable and unfavorable HLA class I alleles and haplotypes in Zambians predominantly infected with clade C human immunodeficiency virus type 1. J. Virol. 2002, 76, 8276–8284. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Los Alamos National Laboratory HIV Molecular Immunoloy Database. 2019. Available online: www.hiv.lanl.gov (accessed on 5 May 2019).
  86. Stewart-Jones, G.B.; Gillespie, G.; Overton, I.M.; Kaul, R.; Roche, P.; McMichael, A.J.; Rowland-Jones, S.; Jones, E.Y. Structures of three HIV-1 HLA-B*5703-peptide complexes and identification of related HLAs potentially associated with long-term nonprogression. J. Immunol. 2005, 175, 2459–2468. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Hansen, S.G.; Ford, J.C.; Lewis, M.S.; Ventura, A.B.; Hughes, C.M.; Coyne-Johnson, L.; Whizin, N.; Oswald, K.; Shoemaker, R.; Swanson, T.; et al. Profound early control of highly pathogenic SIV by an effector memory T-cell vaccine. Nature 2011, 473, 523–527. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Hansen, S.G.; Piatak, M.; Ventura, A.B.; Hughes, C.M.; Gilbride, R.M.; Ford, J.C.; Oswald, K.; Shoemaker, R.; Li, Y.; Lewis, M.S.; et al. Immune clearance of highly pathogenic SIV infection. Nature 2013, 502, 100–104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Hansen, S.G.; Sacha, J.B.; Hughes, C.M.; Ford, J.C.; Burwitz, B.J.; Scholz, I.; Gilbride, R.M.; Lewis, M.S.; Gilliam, A.N.; Ventura, A.B.; et al. Cytomegalovirus vectors violate CD8+ T cell epitope recognition paradigms. Science 2013, 340, 1237874. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Hansen, S.G.; Vieville, C.; Whizin, N.; Coyne-Johnson, L.; Siess, D.C.; Drummond, D.D.; Legasse, A.W.; Axthelm, M.K.; Oswald, K.; Trubey, C.M.; et al. Effector memory T cell responses are associated with protection of rhesus monkeys from mucosal simian immunodeficiency virus challenge. Nat. Med. 2009, 15, 293–299. [Google Scholar] [CrossRef] [Green Version]
  91. Hansen, S.G.; Wu, H.L.; Burwitz, B.J.; Hughes, C.M.; Hammond, K.B.; Ventura, A.B.; Reed, J.S.; Gilbride, R.M.; Ainslie, E.; Morrow, D.W.; et al. Broadly targeted CD8(+) T cell responses restricted by major histocompatibility complex E. Science 2016, 351, 714–720. [Google Scholar] [CrossRef] [Green Version]
  92. Jönsson, P.; Southcombe, J.H.; Santos, A.M.; Huo, J.; Fernandes, R.A.; McColl, J.; Lever, M.; Evans, E.J.; Hudson, A.; Chang, V.T.; et al. Remarkably low affinity of CD4/peptide-major histocompatibility complex class II protein interactions. Proc. Natl. Acad. Sci. USA 2016, 113, 5682–5687. [Google Scholar] [CrossRef] [Green Version]
  93. Phetsouphanh, C.; Aldridge, D.; Marchi, E.; Munier, C.M.L.; Meyerowitz, J.; Murray, L.; Van Vuuren, C.; Goedhals, D.; Fidler, S.; Kelleher, A.; et al. Maintenance of Functional CD57+ Cytolytic CD4+ T Cells in HIV+ Elite Controllers. Front. Immunol. 2019, 10, 1844. [Google Scholar] [CrossRef] [Green Version]
  94. Assarsson, E.; Sidney, J.; Oseroff, C.; Pasquetto, V.; Bui, H.H.; Frahm, N.; Brander, C.; Peters, B.; Grey, H.; Sette, A. A quantitative analysis of the variables affecting the repertoire of T cell specificities recognized after vaccinia virus infection. J. Immunol. 2007, 178, 7890–7901. [Google Scholar] [CrossRef]
  95. Yewdell, J.W. Confronting complexity: Real-world immunodominance in antiviral CD8+ T cell responses. Immunity 2006, 25, 533–543. [Google Scholar] [CrossRef] [Green Version]
  96. Yewdell, J.W. Designing CD8+ T cell vaccines: It’s not rocket science (yet). Curr. Opin. Immunol. 2010, 2, 402–410. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Yewdell, J.W.; Bennink, J.R. Immunodominance in major histocompatibility complex class I-restricted T lymphocyte responses. Annu. Rev. Immunol. 1999, 17, 51–88. [Google Scholar] [CrossRef] [PubMed]
  98. Brehm, M.A.; Pinto, A.K.; Daniels, K.A.; Schneck, J.P.; Welsh, R.M.; Selin, L.K. T cell immunodominance and maintenance of memory regulated by unexpectedly cross-reactive pathogens. Nat. Immunol. 2002, 3, 627–634. [Google Scholar] [CrossRef] [PubMed]
  99. Probst, H.C.; Tschannen, K.; Gallimore, A.; Martinic, M.; Basler, M.; Dumrese, T.; Jones, E.; van den Broek, M.F. Immunodominance of an antiviral cytotoxic T cell response is shaped by the kinetics of viral protein expression. J. Immunol. 2003, 171, 5415–5422. [Google Scholar] [CrossRef] [Green Version]
  100. Van der Most, R.G.; Concepcion, R.J.; Oseroff, C.; Alexander, J.; Southwood, S.; Sidney, J.; Chesnut, R.W.; Ahmed, R.; Sette, A. Uncovering subdominant cytotoxic T-lymphocyte responses in lymphocytic choriomeningitis virus-infected BALB/c mice. J. Virol. 1997, 71, 5110–5114. [Google Scholar] [CrossRef] [Green Version]
  101. Van der Most, R.G.; Murali-Krishna, K.; Lanier, J.G.; Wherry, E.J.; Puglielli, M.T.; Blattman, J.N.; Sette, A.; Ahmed, R. Changing immunodominance patterns in antiviral CD8 T-cell responses after loss of epitope presentation or chronic antigenic stimulation. Virology 2003, 315, 93–102. [Google Scholar] [CrossRef] [Green Version]
  102. Frahm, N.; Yusim, K.; Suscovich, T.J.; Adams, S.; Sidney, J.; Hraber, P.; Hewitt, H.S.; Linde, C.H.; Kavanagh, D.G.; Woodberry, T.; et al. Extensive HLA class I allele promiscuity among viral CTL epitopes. Eur. J. Immunol. 2007, 37, 2419–2433. [Google Scholar] [CrossRef] [Green Version]
  103. Douek, D.C.; Betts, M.R.; Brenchley, J.M.; Hill, B.J.; Ambrozak, D.R.; Ngai, K.L.; Karandikar, N.J.; Casazza, J.P.; Koup, R.A. A novel approach to the analysis of specificity, clonality, and frequency of HIV-specific T cell responses reveals a potential mechanism for control of viral escape. J. Immunol. 2002, 168, 3099–3104. [Google Scholar] [CrossRef]
  104. Barouch, D.H.; O’Brien, K.L.; Simmons, N.L.; King, S.L.; Abbink, P.; Maxfield, L.F.; Sun, Y.H.; La Porte, A.; Riggs, A.M.; Lynch, D.M.; et al. Mosaic HIV-1 vaccines expand the breadth and depth of cellular immune responses in rhesus monkeys. Nat. Med. 2010, 16, 319–323. [Google Scholar] [CrossRef] [Green Version]
  105. Fischer, W.; Perkins, S.; Theiler, J.; Bhattacharya, T.; Yusim, K.; Funkhouser, R.; Kuiken, C.; Haynes, B.; Letvin, N.L.; Walker, B.D.; et al. Polyvalent vaccines for optimal coverage of potential T-cell epitopes in global HIV-1 variants. Nat. Med. 2007, 13, 100–106. [Google Scholar] [CrossRef] [PubMed]
  106. Theiler, J.; Yoon, H.; Yusim, K.; Picker, L.J.; Fruh, K.; Korber, B. Epigraph: A Vaccine Design Tool Applied to an HIV Therapeutic Vaccine and a Pan-Filovirus Vaccine. Sci. Rep. 2016, 6, 33987. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. The HIV-CORE 004 trial: Vaccine immunogens and regimens. (a) Schematic representation of the HIVconsv immunogen derived from 14 conserved regions of the human immunodeficiency virus type 1 (HIV-1) proteome. For each segment, the clade of consensus aa sequence is shown above and the HIV-1 proteins from which it was derived are color-coded. C-terminal CD8+ T-cell and monoclonal antibody epitopes (tags) were added to facilitate preclinical vaccine development and manufacture. The GRIN (Gag-RT-Integrase-Nef) immunogen matched the HIVconsv sequence in 604 out of 776 aa (78%) and these 604 aa of GRIN had 97.6% homology with the HIVconsv protein. The two immunogens are drawn approximately to scale. (b) Three vaccination regimens tested in this trial employed A for Ad35-GRIN—replication-deficient engineered human adenovirus serotype 35 expressing GRIN at 5 × 1010 virus particles; M for MVA.HIVconsv—replication-deficient poxvirus expressing HIVconsv at 2 × 108 plaque-forming units; and D or De for SG2.HIVconsv—”naked” plasmid DNA expressing HIVconsv at 4 mg delivered without or with electroporation, respectively. In vivo intramuscular electroporation was carried out using the TriGrid Delivery System of Ichor Medical Systems, San Diego, CA, USA. Vaccines were delivered by intramuscular needle injection.
Figure 1. The HIV-CORE 004 trial: Vaccine immunogens and regimens. (a) Schematic representation of the HIVconsv immunogen derived from 14 conserved regions of the human immunodeficiency virus type 1 (HIV-1) proteome. For each segment, the clade of consensus aa sequence is shown above and the HIV-1 proteins from which it was derived are color-coded. C-terminal CD8+ T-cell and monoclonal antibody epitopes (tags) were added to facilitate preclinical vaccine development and manufacture. The GRIN (Gag-RT-Integrase-Nef) immunogen matched the HIVconsv sequence in 604 out of 776 aa (78%) and these 604 aa of GRIN had 97.6% homology with the HIVconsv protein. The two immunogens are drawn approximately to scale. (b) Three vaccination regimens tested in this trial employed A for Ad35-GRIN—replication-deficient engineered human adenovirus serotype 35 expressing GRIN at 5 × 1010 virus particles; M for MVA.HIVconsv—replication-deficient poxvirus expressing HIVconsv at 2 × 108 plaque-forming units; and D or De for SG2.HIVconsv—”naked” plasmid DNA expressing HIVconsv at 4 mg delivered without or with electroporation, respectively. In vivo intramuscular electroporation was carried out using the TriGrid Delivery System of Ichor Medical Systems, San Diego, CA, USA. Vaccines were delivered by intramuscular needle injection.
Vaccines 08 00260 g001
Figure 2. Human leukocyte antigen (HLA) class I alleles among Nairobi volunteers. HLA allele (left) and supertypes (right) frequencies are shown (n = 72). U—unclassified supertype.
Figure 2. Human leukocyte antigen (HLA) class I alleles among Nairobi volunteers. HLA allele (left) and supertypes (right) frequencies are shown (n = 72). U—unclassified supertype.
Vaccines 08 00260 g002
Figure 3. Mapping of stimulatory 15-mer peptides. PBMCs were expanded in vitro with 1 of 6 pools of overlapping 15/11 aa peptides spanning the entire HIVconsv immunogen for 10 days to establish short-term cell lines (STCL) and tested in an IFN-γ ELISPOT assay against individual peptides. The frequencies of responding IFN-γ SFU stimulated by peptides are indicated. Peptides and aa shown in red were present in HIVconsv and absent from GRIN. Terminal ‘K’ (blue) were added for solubility (non-HIV). VID—volunteer identification.
Figure 3. Mapping of stimulatory 15-mer peptides. PBMCs were expanded in vitro with 1 of 6 pools of overlapping 15/11 aa peptides spanning the entire HIVconsv immunogen for 10 days to establish short-term cell lines (STCL) and tested in an IFN-γ ELISPOT assay against individual peptides. The frequencies of responding IFN-γ SFU stimulated by peptides are indicated. Peptides and aa shown in red were present in HIVconsv and absent from GRIN. Terminal ‘K’ (blue) were added for solubility (non-HIV). VID—volunteer identification.
Vaccines 08 00260 g003
Figure 4. HC078-expanded STCL derived from volunteer 873 recognized FSVPLDECF (FF9) in the context of three HLA alleles. (a) PBMC of volunteer 873 expanded using HC078 YFSVPLDEGFRKYTA for 10 days showed different IFN-γ staining patterns of CD8+CD3+ cells when stimulated with 721,221 cells stably transfected with HLA-A*03:01, B*15:01 and B*57:03 (right) pulsed with either HC078 or FSVPLDECF peptides (above). Staphylococcal enterotoxin B (SEB) was used as a non-specific positive control. (b) FF9 titrations revealed strong stimulation using HLA-B*57:03 and weak interactions with HLA-A*03:01- and HLA-B*15:01-restricted cells. (c) HC078 STCL was stimulated with HLA-transfected 721.221 cells pulsed with FSVPLDECF (right) and CD8+CD3+ cells were assessed for interaction with either relevant (specific) or irrelevant (non-specific) HLA/peptide tetramers (below) For (a) and (c), the percentage of stained cells is inserted below the gated population.
Figure 4. HC078-expanded STCL derived from volunteer 873 recognized FSVPLDECF (FF9) in the context of three HLA alleles. (a) PBMC of volunteer 873 expanded using HC078 YFSVPLDEGFRKYTA for 10 days showed different IFN-γ staining patterns of CD8+CD3+ cells when stimulated with 721,221 cells stably transfected with HLA-A*03:01, B*15:01 and B*57:03 (right) pulsed with either HC078 or FSVPLDECF peptides (above). Staphylococcal enterotoxin B (SEB) was used as a non-specific positive control. (b) FF9 titrations revealed strong stimulation using HLA-B*57:03 and weak interactions with HLA-A*03:01- and HLA-B*15:01-restricted cells. (c) HC078 STCL was stimulated with HLA-transfected 721.221 cells pulsed with FSVPLDECF (right) and CD8+CD3+ cells were assessed for interaction with either relevant (specific) or irrelevant (non-specific) HLA/peptide tetramers (below) For (a) and (c), the percentage of stained cells is inserted below the gated population.
Vaccines 08 00260 g004
Figure 5. Heterogeneity in HLA-A*02:01-restricted T-cell responses. (a) HLA-A*02:01-positive vaccine recipients were tested for recognition of reportedly HLA-A*02:01-restricted epitopes present in HIVconsv. Their PBMC were expanded with a pool of corresponding parental 15-mers peptides containing the tested epitopes and assessed in an IFN-γ ELISPOT assay. Dotted line indicates the background frequency corresponding to 5 SFU per well. (b) Titration of epitope variants tested for induction of IFN-γ production by STCL.
Figure 5. Heterogeneity in HLA-A*02:01-restricted T-cell responses. (a) HLA-A*02:01-positive vaccine recipients were tested for recognition of reportedly HLA-A*02:01-restricted epitopes present in HIVconsv. Their PBMC were expanded with a pool of corresponding parental 15-mers peptides containing the tested epitopes and assessed in an IFN-γ ELISPOT assay. Dotted line indicates the background frequency corresponding to 5 SFU per well. (b) Titration of epitope variants tested for induction of IFN-γ production by STCL.
Vaccines 08 00260 g005
Figure 6. Mutant epitope prime requires index epitope boost for a functional response. (a) Epitope variants YQYMDDLYV and YQYMAALYV were delivered by the HIVconsv (De for electroporated DNA and M for MVA) and GRIN (Gag-RT-Integrase-Nef) (A for adenovirus) vaccines, respectively. PBMC from volunteer 884 drawn 2, 8, and 12 weeks after DeDeDe, DeDeDeA and DeDeDeAM, respectively, were expanded by either the HIVconsv (black) or GRIN (gray) 15-mer peptide variant and tested against both the parental 15-mer and 9-mer peptides in an ELISPOT assay. (b) Recombinant purified HLA-A*02:01 heavy chain and β2-microglobulin were used to test efficiency of refolding of HLA-peptide complexes in the presence of decreasing test peptide concentrations using a FACS bead assay. (c) Killing of peptide-pulsed HLA-A*02:01 target cells by a YQYMDDLYV-specific CD8+ T cell clone. CQYMDDLYV present in 1.32% of database sequences was used as a variant example of a strong binder with impaired T-cell stimulation.
Figure 6. Mutant epitope prime requires index epitope boost for a functional response. (a) Epitope variants YQYMDDLYV and YQYMAALYV were delivered by the HIVconsv (De for electroporated DNA and M for MVA) and GRIN (Gag-RT-Integrase-Nef) (A for adenovirus) vaccines, respectively. PBMC from volunteer 884 drawn 2, 8, and 12 weeks after DeDeDe, DeDeDeA and DeDeDeAM, respectively, were expanded by either the HIVconsv (black) or GRIN (gray) 15-mer peptide variant and tested against both the parental 15-mer and 9-mer peptides in an ELISPOT assay. (b) Recombinant purified HLA-A*02:01 heavy chain and β2-microglobulin were used to test efficiency of refolding of HLA-peptide complexes in the presence of decreasing test peptide concentrations using a FACS bead assay. (c) Killing of peptide-pulsed HLA-A*02:01 target cells by a YQYMDDLYV-specific CD8+ T cell clone. CQYMDDLYV present in 1.32% of database sequences was used as a variant example of a strong binder with impaired T-cell stimulation.
Vaccines 08 00260 g006
Table 1. Summary of stimulatory human leukocyte antigen (HLA) class I T-cell epitopes in the Nairobi cohort.
Table 1. Summary of stimulatory human leukocyte antigen (HLA) class I T-cell epitopes in the Nairobi cohort.
No.ParentalVID 2NameShorterReportedPredicted 1Confirmed
SequenceSequenceHLAHLAHLA
HC014EWDRIYKRWIILGLN 829YN10YKRWIILGLNNot rep’dB*27:03
HC049(K)NFPISPIETVPVKLK831SL10SPIETVPVKLB*81:01B*81:01
IL9IETVPVKL B*40:01
B*53:01
HC078YFSVPLDEGFRKYTA873FF9FSVPLDEGFB*57:03B*57:03B*57:03
B*15:01B*15:01
A*03:01A*03:01
A*26:01
HC088GSPAIFQSSMTKILE830AK9AIFQSSMTKA*03:01A*03:01A*03:01
A*11:01
II9IFQSSMTKINot rep’d B*51:01
866SK11SPAIFQSSMTKA*11:01 B*07:05 ♣
A*11:01
B*53:01 ♣
GM9GSPAIFQSSMNot rep’dB*07:05
SM9SPAIFQSSMB7B*07:05B*07:05 ♣
B*53:01B*53:01 ♣
C*04:02
HC164VQMAVFIHNFKRKGGI830AR9AVFIHNFKRA*03:01A*03:01
A*74:01
MR10 MAVFIHNFKR Not rep’d A*03:01
A*66:01
A*68:01
MK9MAVFIHNFKA*03:01A*03:01
A*74:01
B*51:01
VR8 VFIHNFKR Not rep’d A*66:01
A*68:01
889MR10MAVFIHNFKRNot rep’d
MK9MAVFIHNFKNot rep’dC*02:02
VR8VFIHNFKRNot rep’d
HC176VVPRRKAKIIRDYGK831KK10KAKIIRDYGKNot rep’d
889KK8KIIRDYGKNot rep’d
KY8KAKIIRDYNot rep’d
HC177RKAKIIRDYGKQMAG810RK11RKAKIIRDYGKNot rep’d
RY9RKAKIIRDYB*15:03B*15:03
1 Epitopes were predicted using the LANL-HMID Immune Epitope Database (IEDB). 2 VID—volunteer identification. Please refer to Table S1 for volunteers’ HLAs. ♣—Previously unreported candidates for LANL-HMID “A” list.

Share and Cite

MDPI and ACS Style

Mohamed, Y.S.; Borthwick, N.J.; Moyo, N.; Murakoshi, H.; Akahoshi, T.; Siliquini, F.; Hannoun, Z.; Crook, A.; Hayes, P.; Fast, P.E.; et al. Specificity of CD8+ T-Cell Responses Following Vaccination with Conserved Regions of HIV-1 in Nairobi, Kenya. Vaccines 2020, 8, 260. https://doi.org/10.3390/vaccines8020260

AMA Style

Mohamed YS, Borthwick NJ, Moyo N, Murakoshi H, Akahoshi T, Siliquini F, Hannoun Z, Crook A, Hayes P, Fast PE, et al. Specificity of CD8+ T-Cell Responses Following Vaccination with Conserved Regions of HIV-1 in Nairobi, Kenya. Vaccines. 2020; 8(2):260. https://doi.org/10.3390/vaccines8020260

Chicago/Turabian Style

Mohamed, Yehia S., Nicola J. Borthwick, Nathifa Moyo, Hayato Murakoshi, Tomohiro Akahoshi, Francesca Siliquini, Zara Hannoun, Alison Crook, Peter Hayes, Patricia E. Fast, and et al. 2020. "Specificity of CD8+ T-Cell Responses Following Vaccination with Conserved Regions of HIV-1 in Nairobi, Kenya" Vaccines 8, no. 2: 260. https://doi.org/10.3390/vaccines8020260

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