Next Article in Journal
Characterization of Immune Response towards Generation of Universal Anti-HA-Stalk Antibodies after Immunization of Broiler Hens with Triple H5N1/NA-HA-M1 VLPs
Previous Article in Journal
Characterization of H9N2 Avian Influenza Viruses Isolated from Poultry Products in a Mouse Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Pre-Treatment Integrase Inhibitor Resistance and Natural Polymorphisms among HIV-1 Subtype C Infected Patients in Ethiopia

by
Dawit Assefa Arimide
1,2,*,
Zsófia Ilona Szojka
3,
Kidist Zealiyas
2,†,
Atsbeha Gebreegziabxier
2,†,
Fekadu Adugna
4,
Sviataslau Sasinovich
1,
Per Björkman
1,5 and
Patrik Medstrand
1,6
1
Department of Translational Medicine, Lund University, 205 02 Malmo, Sweden
2
TB/HIV Department, Ethiopian Public Health Institute, Addis Ababa 1242, Ethiopia
3
Department of Laboratory Medicine, Lund University, 222 42 Lund, Sweden
4
NPO–HIV/AIDS, World Health Organization, Addis Ababa 3069, Ethiopia
5
Department of Infectious Diseases, Skane University Hospital, 205 02 Malmo, Sweden
6
Clinical Microbiology and Infectious Disease Control, Skane University Hospital, 221 85 Lund, Sweden
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Viruses 2022, 14(4), 729; https://doi.org/10.3390/v14040729
Submission received: 4 March 2022 / Revised: 18 March 2022 / Accepted: 28 March 2022 / Published: 30 March 2022
(This article belongs to the Section Human Virology and Viral Diseases)

Abstract

:
Dolutegravir-based antiretroviral therapy (ART) has been scaled up in many developing countries, including Ethiopia. However, subtype-dependent polymorphic differences might influence the occurrence of HIV-drug-resistance mutations (HIVDRMs). We analyzed the prevalence of pre-treatment integrase strand transfer inhibitor (INSTI) HIVDRMs and naturally occurring polymorphisms (NOPs) of the integrase gene, using plasma samples collected as part of the national HIVDR survey in Ethiopia in 2017. We included a total of 460 HIV-1 integrase gene sequences from INSTI-naïve (n = 373 ART-naïve and n = 87 ART-experienced) patients. No dolutegravir-associated HIVDRMs were detected, regardless of previous exposure to ART. However, we found E92G in one ART-naïve patient specimen and accessory mutations in 20/460 (4.3%) of the specimens. Moreover, among the 288 integrase amino acid positions of the subtype C, 187/288 (64.9%) were conserved (<1.0% variability). Analysis of the genetic barrier showed that the Q148H/K/R dolutegravir resistance pathway was less selected in subtype C. Docking analysis of the dolutegravir showed that protease- and reverse-transcriptase-associated HIVDRMs did not affect the native structure of the HIV-1 integrase. Our results support the implementation of a wide scale-up of dolutegravir-based regimes. However, the detection of polymorphisms contributing to INSTI warrants the continuous surveillance of INSTI resistance.

1. Introduction

Following the global increase of pre-treatment drug resistance (PDR) to non-nucleoside reverse transcriptase inhibitors (NNRTIs), the World Health Organization (WHO) recommended the transition from NNRTI to integrase strand transfer inhibitor (INSTI)-based regimens in both treatment-naïve and treatment-experienced patients [1,2,3]. Several low- and middle-income countries have already transitioned to the dolutegravir (DTG)-based regimen, and many more are in the planning phase, so millions of people living with HIV will soon receive DTG combined with two nucleoside reverse transcriptase inhibitors (NRTIs) as first- and second-line therapies [2,4].
HIV-1 integrase (IN), which comprises 288 amino acids encoded by the 5′-end of the HIV pol (polymerase) gene, plays a vital role in HIV-1 replication by catalyzing two distinct reactions: 3′-end processing and strand transfer [5,6,7]. IN consists of three functional domains: the N-terminal domain (NTD) (aa: 1–50), which contains a highly conserved histidine–histidine–cysteine–cysteine (H12H16C40C43) motif that coordinates zinc binding and favors multimerization of the IN subunit [8]; the catalytic core domain (CCD) (aa: 51–212), which contains the catalytic triad D64D116E152 (known as the DDE motif) that plays an essential role in IN enzymatic activity; and the C-terminal domain (CTD) (aa: 213–288), which is involved in binding to viral and cellular DNA, and in protein oligomerization and interactions with the reverse transcriptase [5,6,7,9].
INSTIs inhibit the HIV-1 integrase strand transfer steps to block the integration of HIV viral DNA into the host cell chromosomal DNA through competitive binding to the enzyme’s active site [7,10]. There are currently five US Food and Drug Administration (FDA)-approved drugs belonging to this therapeutic class: raltegravir (RAL), elvitegravir (EVG), DTG, bictegravir (BIC), and cabotegravir (CAB) [11]. RAL and EVG were the first-generation INSTIs to be used clinically; however, their relatively low genetic barrier for resistance and the extensive cross-resistance between them limit their efficiency [12,13]. DTG and BIC are second-generation INSTIs shown to be highly effective in both treatment-naive and treatment-experienced individuals with good tolerability and a high genetic barrier to resistance [12,14]. Pooled analysis of resistance data conducted by Yang et al. (2019) indicated that the development of resistance to DTG and BIC was rare [12]. However, with the wide scale-up of DTG, gradual development and transmission of HIVDR against INSTIs will be inevitable and can render existing therapies ineffective, thereby increasing the risk of virological failure, disease progression, and mortality [12,15,16,17,18,19,20,21,22,23,24,25,26,27,28].
Although non-B subtypes dominate the global HIV epidemic, most clinical and virological studies on DTG were based on subtype B. However, subtype-dependent differences in naturally occurring polymorphisms (NOPs) have been implicated in the development of different mutational pathways, leading to varying levels of drug resistance against INSTIs among different HIV-1 subtypes [5,13,29,30,31,32,33,34]. Q148H and G140S, which confer resistance to RAL and EVG and cross-resistance to DTG, appear more frequently in subtype B than in non-B subtypes [31]. Similarly, R263K is mainly present in subtype B, while G118R has a pathway in selecting DTG resistance in non-B subtype viruses [13,22,35,36].
HIV-1 sequences and structure-based analyses also showed that subtype-specific NOPs, especially at the active site of IN, can affect the genetic barrier to drug resistance by influencing the selection of resistance mutations, native protein structure, and the function of the drug-mediated inhibition of the enzyme [29,30,32,33,37].
In 2019, an estimated 669,236 people were living with HIV in Ethiopia, and the epidemic was dominated by subtype C [38,39]. Similar to many other countries in sub-Saharan Africa, Ethiopia has implemented the test-and-treat strategy, with DTG-based regimens recommended as the first-line antiretroviral therapy (ART) [40]. However, there is limited knowledge of the frequency and characteristics of NOPs of IN or their effect on the development of INSTI resistance. This study aimed to investigate HIV-1 IN genotypic profile to evaluate the prevalence of pre-treatment DRMs and NOPs that might affect the genetic barrier to the emergence of resistance in INTSI-naïve patients in Ethiopia infected with HIV-1 subtype C.

2. Materials and Methods

2.1. Study Design

In this study, we used plasma samples collected from HIV-1-infected patients as part of a national HIVDR survey conducted in Ethiopia. A cross-sectional survey was conducted in 2017 among treatment-naïve patients and patients on first- and second-line regimens in selected health facilities from different parts of the country according to the WHO-recommended HIVDR survey [41]. After obtaining written informed consent from each participant, 10 mL of blood was collected by venipuncture for CD4+ T-cell count, viral load, and HIVDR genotyping. Basic demographic and clinical information were also collected during the survey using a standardized questionnaire. Specimens were transported to the Ethiopian Public Health Institute (EPHI) on dry ice for viral load testing and long-term storage at −80 °C. HIV-1 VL was determined using the Abbott RealTime HIV-1 assay (Abbott Molecular Inc., Des Plaines, IL, USA). Using 1000 copies/mL as a viral load suppression threshold based on the WHO recommendation [42], all samples with a viral load ≥1000 copies/mL were then shipped to the National Institute of Respiratory Diseases-Mexico (INER) laboratory for HIVDR genotyping.

2.2. HIV-1 Genotyping

Genotyping of the integrase region was performed using an in-house-developed and -validated protocol for IN [43]. Amplicons obtained by the nested PCR method were used for Sanger sequencing using the BigDye technology on the ABI Prism 3730 Genetic Analyzer (Applied Biosystems, Foster City, CA, USA). Sequence assembly and editing were performed using the RECall V 2.0 HIV-1 sequencing analysis tool (University of British Columbia, Vancouver, BC, Canada) [44]. Sequence quality control was performed using the WHO tool (https://sequenceqc-dev.bccfe.ca/who_qc (accessed on 28 June 2021)) and the Quality Control program of the Los Alamos HIV sequence database (https://www.hiv.lanl.gov (accessed on 28 June 2021)).

2.3. Subtype Determination Using HIV-1 Integrase Sequences

The HIV-1 subtyping was performed using the online automated subtyping tools REGA v3.0 [45], COMET [46], and the jumping profile Hidden Markov Model (jpHMM) [47]. Subtyping was further confirmed by Maximum likelihood (ML) phylogenetic tree analysis with the IN references sequences from HIV-1 subtype (A-K) and recombinant virus downloaded from the Los Alamos database (http://www.hiv.lanl.gov (accessed on 3 July 2021)). Multiple sequence alignment was conducted using MAFFT version 7 [48] and was then manually edited using BioEdit V7.0.9.0 [49,50] until a perfect codon alignment was obtained. ML tree topology was constructed using the online version of PhyML v 3.0 [51] with the GTR+I+Γ nucleotide-substitution model (using the estimated proportion of invariable sites and four gamma categories). A heuristic tree search was performed using the SPR branch-swapping algorithm. Branch support was determined with aLRT-SH (approximate likelihood ratio test, Shimodaira Hasegawa-like) [52]. Clusters were defined as monophyletic clades with aLRT-SH support ≥0.9. The subtype-resolved ML phylogeny trees were visualized using the FigTree v1.4.0 program. Sequence(s) that formed a cluster with the reference sequences belonging to the same subtype were assigned to that subtype.

2.4. HIV-1 Drug Resistance Analysis

INSTI-associated mutations were identified using the Stanford HIV Drug Resistance Database (HIVdB v9.0) (https://hivdb.stanford.edu/hivdb/by-mutations (accessed on 7 July 2021)). INSTI DRMs were categorized as major resistance mutations, accessory resistance mutations, and other mutations according to the Stanford HIV Drug Resistance Database. Major resistance mutations were primarily nonpolymorphic DRMs that caused a significant reduction in INSTI susceptibility, even when they occurred alone. Accessory mutations were nonpolymorphic or minimally polymorphic mutations that caused only low-level reduction of INSTI susceptibility when they occurred alone, but may have augmented resistance and/or restored the fitness of viral mutants with major resistance mutations. The other mutations included highly polymorphic and/or rare nonpolymorphic mutations that may have been weakly associated (uncertain role) with drug resistance. We further extensively investigated all amino acid positions associated with decreased INSTI susceptibility. Samples harboring resistant and/or a mixture of wild-type and resistant amino acids were considered resistant.

2.5. HIV-1 Subtype C Integrase Polymorphism and Conservation Analysis

For this analysis, only HIV-1 subtype C sequences were used. Briefly, multiple sequence alignment was conducted using MAFFT version 7 [48] and was then manually edited using BioEdit V7.0.9.0 [49,50] until a perfect codon alignment was obtained. The nucleotide sequences were translated to an amino acid sequence. Then, each amino acid along the 288 IN positions was extensively investigated for the presence of primary mutations and of nonpolymorphic and polymorphic mutations associated with resistance to INSTI. The prevalence of each amino acid at each IN position was determined and compared to the HIV-1 subtype B reference sequence (GenBank accession number: K03455). We defined NOPs as substitutions within the HIV-1 IN that occurred in ≥1% of the sequences for this analysis [6]. The positions with ≥20% substitutions were defined as highly polymorphic, while those with ≤0.5% variability were considered highly conserved.

2.6. Generation of Consensus HIV-1 Integrase Sequence

To comprehensively describe the variability (polymorphism) in the IN sequences, we downloaded global subtype B and C IN sequences that matched the region (HXB2: 4230- 5093 relative to HXB2 clone) from the HIV Los Alamos National Library (LANL) database (https://www.hiv.lanl.gov (accessed on 13 July 2021)). To avoid the overestimation of variant calling and ensure the sequences included in the analysis were from INSTI-naïve patients, only sequences before 2007 (before the FDA approved INSTIs) were used. The quality of all HIV-1 sequences was verified using the online Quality Control program (http://www.hiv.lanl.gov (accessed on 14 July 2021)). Sequences with stop codons and/or frameshifts and/or poor quality were removed from the analysis. Only one sequence per patient was retained. For a patient with multiple sequences, the earliest sequence was selected and used. The consensus amino acid sequence for IN was generated for Ethiopian HIV-1 subtype C, the global HIV-1 subtype B, and the global subtype C sequence using BioEdit V7.0.9.0 [49,50]. For positions where two amino acids occurred at frequencies higher than 30%, both amino acids were represented, and the first letter seen at the consensus represented the most prevalent amino acid.
Furthermore, to assess the impact of previous exposure to ART on IN gene NOPs, the consensus amino acid sequences of IN from the ART-naïve and ART-experienced patients were generated and compared. Similarly, we also compared the consensus amino acid sequences of IN from patients with one or more major HIVDRMs to protease inhibitor (PI), NRTI, and/or NNRTIs (HIVDR group) with those with no major HIVDRMs (no-HIVDR groups) in their corresponding protease/reverse transcriptase (PR/RT) gene.

2.7. Genetic Barrier to Integrase Strand-Transfer Inhibitor Resistance

To assess differences in the genetic barrier for evolution of drug-resistance substitutions between subtypes C and subtype B, we compared Ethiopian HIV-1 subtype C IN sequences obtained from INSTI-naïve patients and global HIV-1 subtype B sequences obtained from LANL (INSTI-naïve, collected before 2007). We calculated the genetic barrier to INSTI resistance for 10 major INSTI resistance amino acid positions (19 substitutions) using a previously published method [53]. Briefly, we first determined the extent of natural diversity at each selected position in our dataset of Ethiopian HIV-1 subtype C IN sequences and global subtype B IN sequences by identifying all wild-type triplets and their prevalence. Next, we compute genetic barrier score for each wild-type triplet to evolve to resistant amino acid at the specific selected position. The genetic barrier was calculated as the sum of transitions and/or transversions required to evolve to any major drug-resistance substitution. We used a score of 1 for transition (A↔G and C↔T), 2.5 for transversion (A↔C, A↔T, G↔C, G↔T), and 0 when no change was needed, as described by Nguyen et al. (2012) [53]. The smallest number (minimal score) of transversion and/or transition required for evolution from wild-type codon to resistant codon were used to calculate the genetic barrier.

2.8. Modeling and In Silico Predictions of HIV-1 Integrase and Dolutegravir Interaction

For in silico predictions, 20 randomly selected (10 from each ART-naïve (PDR) and ART-experienced (ADR)) sequences were used. The ART-naïve IN sequences used in our analysis had no HIVDRMs against NRTI, NNRTI, and/or PI in their corresponding PR/RT gene, while the ART-experienced group had one or more HIVDRMs against NRTI, NNRTI, and/or PI. A multiple-sequence alignment of amino acid sequences (without any gap) was made using ClustalW (https://www.genome.jp/tools-bin/clustalw (accessed on 1 November 2021)). An amino acid identity matrix was created with Clustal 12.1 (https://www.ebi.ac.uk/Tools/msa/clustalo (accessed on 1 November 2021)) and visualized using GraphPad Prism 8.
The crystallographic structure of full-length HIV-1 IN (accession number: 6u8q.pdb) was obtained from the Protein Data Bank (www.rcsb.org (accessed on 2 November 2021)) [54]. To visualize both the PDR and ADR HIV-1 IN, the 6u8q was modified by using UCSF-Chimera at 12 amino acid positions (see Table S2), and a monomer was used in the docking prediction. The structure (6u8q) originally included a DNA fragment and DTG. After removing all ligands, the DNA fragment and water molecules from the crystal structure, receptor, and ligand–DTG files were separately saved for further analysis. MGL Tools (Version 1.5.7rc1) was used for creating .pdbqt files of the receptor and ligands needed for docking with Autodock Vina (Vina) (Version 1.1.2) [55,56]. Ligands were docked to the binding site cavity using x = 211, 63 Å; y = 205, 453 Å; and z = 171, 895 Å Cartesian coordinates that used the catalytic site in the monomer of HIV-1 IN. The grid box dimensions used for the search space were 50 Å × 40 Å × 40 Å. Docking calculations were performed with an exhaustiveness option of 8 (average accuracy) and an energy range of 3. Validation of the docking method was performed by redocking DTG to the modified crystal structure to the modified above-mentioned structure.

2.9. Statistical Analysis

Fisher’s exact test, the Chi-squared test, and the Mann–Whitney U-test were used to evaluate the statistical differences between groups. p-values ≤ 0.05 were considered statistically significant.

3. Results

A total of 460 IN sequences obtained from INSTI-naïve patients were included in the analysis. Among these, 373 sequences were from patients who did not report exposure to any antiretroviral drug at the time of specimen collection (ART-naive), while 87 sequences were from ART-experienced (NNRT-based or PI-based regimens) patients, with virological failure (viral load ≥ 1000 copies/mL) while on a first-line (n = 41) or second-line (n = 46) regimen.

3.1. HIV-1 Subtyping

Online subtyping and the subsequent phylogenetic analysis results showed that 98.5% (453/460) of the sequences were subtype C, while 0.43% (2/460), 0.22% (1/460), 0.22% (1/460), 0.22% (1/460), 0.22% (1/460), and 0.22% (1/460) were subtype B, subtype A1, CRF10_CD, CRF02_AG, CRF49_cpx and CRF_A2D, respectively (Figure 1).
The phylogenetic tree in Figure 1 contains a total of 874 sequences, including Ethiopian sequences (n = 460) and (n = 414) integrase reference sequences for HIV-1 subtypes (A–K) and circulating recombinant forms downloaded from the HIV-1 LANL database. An ML tree was constructed using the online version of PhyML v 3.0. The reference sequences from the Los Alamos National Laboratory are in black in the figure. All the Ethiopian sequence’s clusters with the HIV-1 subtype C reference sequence are in green, while the non-subtype C Ethiopian sequences are in pink.

3.2. Prevalence of Major Integrase Strand-Transfer Inhibitor Resistance Mutations

No major DRMs known to be associated with DTG resistance (T66K, E92Q, G118R, E138K/A/T, G140S/A/C, Q148H/R/K, N155H, or R263K) were detected among INSTI- naïve individuals, regardless of previous exposure to ART. However, one (0.22%) sequence from a person without previous ART exposure was found to harbor E92G, a mutation that moderately reduces EVG susceptibility but does not reduce susceptibility to RAL and DTG.
A total of 4.4% (20/460) of the sequences contained five different IN accessory mutations: −E157Q (2.39%), G163R/K (0.65%), Q95K (0.65%), T97A (0.43%), and G149A (0.22%). There was no significant difference in the prevalence of accessory mutations among ART-naïve and ART-experienced patients (p = 0.9) (Table 1). Only one accessory mutation per sequence was detected, except for one sequence with two (G149A and E157Q) accessory mutations. In addition, other mutations including M50I (18.5%, 85/460), L74I/M (2.8%, 13/460), S119R, (0.9%,4/460), V151I, (1.3%, 6/460), and D230N (0.4%, 2/460) were also detected.

3.3. Integrase Strand-Transfer Inhibitor Resistance among Patents on Antiretroviral Therapy

To assess the impact of ART exposure to NRTI, NNRTI, and/or PI on the selection of INSTI-resistance mutations, we further compared the INSTI HIVDRMs from patients with one or more major HIVDR mutations to NRTI, NNRTI, and/or PI (HIVDR group) with those with no HIVDRMs in their corresponding PR/RT genes (no-HIVDR group) (Figure 2).
Briefly, among the total 460 IN sequences used in our analysis, 327 had a corresponding PR/RT gene sequence, of which 234 had no major HIVDRMs (no-HIVDR group), while 93 of the sequences (HIVDR group) had one or more HIVDRMs against the NRTI, NNRTI, and/or PI (see Table S1). No major INSTI HIVDRMs were detected in either of these groups, and there was no significant difference in the presence of accessory mutations with regard to previous ART exposure, nor with regard to DRMs toward other ARVs. Among the HIVDR and no-HIVDR groups, 3.2% (3/93) and 4.7% (11/234) accessory mutations were detected, respectively (p = 0.8); while 4.29% (15/373) and 5.75% (5/87) accessory mutations were detected among ART-naïve and ART-experienced groups, respectively (p = 0.6).
High similarity was also observed when comparing the consensus sequence from ART-naïve and ART-experienced patients, as shown in Figure 3. Similarly, our comparison of the consensus sequences from the HIVDR and non-HIVDR groups also showed high similarity between the two consensus sequences, except at positions K215N, T218L, and R269, where the HIVDR group had one amino acid; while the no-HIVDR group had a mixture of amino acids at positions T215K/N, T218I/L, and R269R/K, respectively (Figure 3).

3.4. Prevalence of Naturally Occurring Integrase Polymorphisms in HIV-1 Subtype C

An alignment of the 453 HIV-1 subtype C IN sequences from the INSTI-naïve Ethiopian patients was extensively analyzed and compared to the HIV-1 subtype B reference sequence (GenBank accession number: K03455). Based on our definition of polymorphism (≥1.0% variability), an overall 64.9% (187/288) amino acid positions of the IN were conserved. The conservations of the NTD, CCD, and CTD were 60% (30/60), 66.1% (107/162), and 66.8% (50/76), respectively. The distribution of polymorphisms in 453 HIV-1 subtype C IN sequences is shown in Figure 4.

3.5. Analysis of the N-Terminal Domain (NTD)

Within the NTD, the Zn-binding motif (H12H16C40C43) involved in the multimerization of the IN subunit, stabilization of folding, and interaction with LEDGF/p75 were highly conserved [6]. However, amino acid positions, D10E, S24N, D25E, V31I, and M50I were highly polymorphic (>20.0% variability). We also observed that the residue E10 had been replaced by D (aspartic acid) in 97.8% of sequences, which might be the signature of subtype C (Figure 4).

3.6. Analysis of the Catalytic Core Domain (CCD)

In the CCD, the catalytic triad D64D116E152 was highly conserved, and was found within the conserved regions 61–70, 114–118, and 152–155, respectively. The critical positions for DNA-binding HIV-1 integration and replication (Q62, H67, N120, N144, Q148, and N155) [57] and the residue involved in the chemical bond and hydrophobic contact with the LEDGF/p75 [6] (A128-A129-W131-W132-Q168-E170-T174-M178) were also highly conserved. However, amino acids at codon positions G163, V165, D167, H171, and K173 within the I161-K173 region known to be involved in the noncanonical nuclear localization signal [6,22], and the K188 within the KRK motif (K186, R187, K188), which is vital for the integrase:integrase:oligomerization at the dimer:dimer interface [6,22], showed 28.5% variability.
Among the INSTI-mutation positions in the CCD residues that directly reduced the INTSI susceptibility, H51, T66, E92, F121, G140, Y143, Q146, S147, Q148, S153, N155, and E157Q were highly conserved, except for codon position E157Q, which was a polymorphic position (>1.0% variability). However, a highly polymorphic residue in the CCD including V72I, I84M, F100Y, L101I, T112V, T124A, T125A, R127K, K136Q, D167E, K188R, and V201I was observed.

3.7. Analysis of the C-Terminal Domain (CTD)

Within CTD, the two large consecutive residues, L241-Q252 and I257-K264, which are involved in the binding of viral and cellular DNA, were found to be highly conserved, except for positions I251 and V257, which were mutated to I251L and V259I in 3.5% and 0.7% of the sequences, respectively. However, the important positions for DNA binding and integrase multimerization (K258, V260, R262, R263, and K264) [6] were fully conserved.
Our analysis also showed that 24 amino acid positions were highly polymorphic (>20.0% variability): D10E, K14R, S24N, D25E, V31I, M50I, V72I, I84M, F100Y, L101I, T112V, T124A, T125A, R127K, K136Q, D167E, K188R, V201I, K215N, T218I, A265V, R269K, D278A, and S283G. Six of these (D10E, K14R, S24N, D25E, V31I, and M50I) belonged to the NTD, whereas 12 (V72I, I84M, F100Y, L101I, T112V, T124A, T125A, R127K, K136Q, D167E, K188R, and V201I) belonged to the CCD, and the other 6 (K215N, T218I, A265V, R269K, D278A, and S283G) belonged to the CTD.
Our comparison of the NOPs’ distribution with the global subtype B and global subtype C sequences downloaded from LANL showed that the Ethiopian HIV-1 subtype C IN sequences had a high similarity to the global subtype C sequence, but were quite different from the global subtype B, as shown in Figure 5.

3.8. Analysis of the Subtype Consensus Integrase Sequences

The consensus IN sequence for the global HIV-1 subtype B and global subtype C were generated using 1884 and 1410 sequences, respectively. Our comparison of the 288 amino acid sequence alignment of the consensus Ethiopian HIV-1 subtype C with the global HIV-1 subtype C showed high similarity, except for the mixture of amino acid sequences at positions 25E/D, 100Y/F,124T/A, 136K/Q, 167E/D, 215K/N, and 218I/L in the Ethiopian consensus; and 50M/I, 72I/V, and 265A/V in the global HIV-1 subtype C consensus sequence. However, it differed from the global subtype B consensus at eight positions with complete amino acid replacement (31, 112, 125, 201, 218, 234, 278, 283), while a mixture of amino acids was detected at positions 11E/D, 72I/V, and 101I/L in the global subtype B consensus sequences and at 24N/S, 25E/D, 100Y/F, 124T/A, 136K/Q, 167E/D, 215K/N, and 269K/R in the Ethiopian subtype C consensus sequence (Figure 6).

3.9. Genetic Barrier to Dolutegravir Resistance

In this study,19 substitutions conferring major resistance to DTG at 10 amino acid positions in the IN (T66A/I/K, E92G, G118R, E138K/A/T, G140S/A/C, Y143R/C/H, S147G, Q148H/R/K, N155H, and R263K) were assessed to explore the genetic barrier to DTG. For each codon, the number of transitions and/or transversions required for a IN drug resistance associated substitution were calculated. A total of 1884 global HIV-1 subtype B sequences and 453 Ethiopian subtype C sequences from INSTI-naïve patients were compared for differences in the genetic barrier to INSTI resistance (Table 2).
Overall, the sequence analysis of the two subtypes showed similar predominant codon use at the selected amino acid positions, resulting in a similar minimum score for the genetic barrier to DTG. However, at position 140, the predominant codons in subtype C were GGG (53.6%) and GGA (45.9%). In contrast, in subtype B, GGC (85.0%) was the predominant codon resulting in a difference in the calculated genetic barrier at this position. For subtype C, two transversions (minimum score of 5) were required to mutate to G140C (GGG/A to ATG/C); while for subtype B, one transversion and transition (minimum score: 3.5) were required to mutate to G140C (GGC to TGT). Similarly, a two-point mutation (one transversion and one transition) (minimum score of 3.5) was required to mutate to G140S (GGG/A to AGT/C) for subtype C; while subtype B required a one-step transition (minimum score of 1) (GGC to AGC).

3.10. Impact of Protease and Reverse-Transcriptase Drug-Resistance Mutation on the Structure of HIV-1 Integrase

The effects of HIVDRMs in HIV-1 PR and/or RT on the secondary structure of HIV-1 IN were investigated on 20 sequences: 10 from ART-naïve (PDR) and 10 from ART-experienced (ADR) individuals representative of randomly selected HIV-1 IN sequences. The sequence identity matrix (Figure 7a) showed that all the sequences were more than 92% identical at the amino acid level, and there were no major differences between the two main groups. To study the effects of PR and RT drug-induced resistance on the structure of HIV-1 IN, chain A of the 6u8q structure was modified at 12 positions to represent both the ADR and the PDR sequences (see Table S2). The alignment of the monomers of the PDR and ADR INs did not result in any differences between the two groups. DTG was successfully docked to both the PDR and ADR IN by Autodock Vina (Figure 7c), and the docking score was −6.5 kcal/mol, which was at a similar position as the original DTG ligand.

4. Discussion

Overall, our results revealed no major DTG associated HIVDRM mutations among INSTI-naïve individuals, regardless of previous exposure to ART. In one individual, the E92EG mutation was found, which moderately reduced EVG susceptibility, but had no effect on DTG. However, INSTI accessory mutations and NOPs, which could influence INSTI susceptibility and the genetic barrier to INSTI resistance, were detected. Our polymorphism analysis showed that 64.9% (187/288) of amino acid positions of the HIV-1 subtype C IN sequences from INSTI-naïve individuals were conserved (<1.0% variability). The majority of amino acids involved in key functions of the enzyme (the HHCC motif and the DDE motifs [6,22]) were fully conserved. The genetic barriers to DTG resistance were similar at selected amino acid positions for subtypes B and C, except that subtype C had a higher genetic barrier for the G140C and G140S mutations, highlighting that the Q148H/K/R DTG resistance pathway was selected less in subtype C. Docking analysis of the DTG showed that the PR- and RT-associated HIVDRM did not affect the structure of the HIV-1 IN, supporting the use of DTG as a salvage therapy for patients with resistance to drugs targeting these enzymes.
The absence of major INSTI DRMs among INSTI-naïve patients in our study was consistent with other studies from Africa [58,59,60,61,62,63,64], Asia [65,66,67], and Europe [68,69,70], showing no or highly infrequent major INSTI mutations among INSTI-naïve patients. Our finding was not unexpected, and was in line with studies from other settings based on samples obtained before the rollout of DTG [71,72,73]. However, following the wide scale-up of DTG, an increase in DTG resistance has been reported, especially in persons receiving DTG monotherapy [15,19,23,24,25,26,27,28]. Hitherto, the prevalence of transmitted resistance to DTG resistance has been low [20,21,22,23]. Similarly, in Ethiopia, after implementing the test-and-treat strategy, an increased number of patients will be on a DTG-based regimen. Thus, the emergence of INSTI resistance is expected, especially in settings with low access to viral load monitoring, delaying the identification of patients with treatment failure and increasing the risk of HIV drug resistance [74].
When present alone, accessory mutations have a minimal effect on INSTI susceptibility, but may serve to augment resistance and/or restore the fitness of viral mutants with major resistance mutations [5,30]. INSTI accessory mutations were detected in 20 (4.4%) of our specimens, and were equally distributed in both ART-naive and ART-experienced patients. Similar to our findings, different studies [67,72,73,75] revealed that NOPs were common among INSTI-naive patients. However, the prevalence differed with HIV-1 subtypes or circulating recombinant forms.
E157Q was the most common nonpolymorphic accessory mutation detected in our analysis. It is a natural polymorphism present in 1–10% of untreated individuals, depending on the subtype. It has no effect on the susceptibility of INSTI. However, it may act as a compensatory substitution for R263K-induced resistance to DTG [76]. Q95K was among the other nonpolymorphic accessory INSTI resistance detected in our study, and it had little, if any, effect on drug susceptibility to INSTI; however, in the presence of a N155H mutation, it increased INSTI resistance and improved the impaired replication of the virus [77].
L74M/I (2.9%) and M50I (18.8%) were the other polymorphic mutations detected in our study. L74M/I has been reported at levels between 0.5–20% in the untreated population, with a high prevalence in subtypes A, G, and A/G recombinants. It does not decrease INSTI susceptibility alone, but it can contribute to a high-level resistance when occurring with major INSTI-resistance mutations, mainly the Q148H/K/R mutation [24,58,78,79]. Studies in South Africa, Brazil and Europe have also confirmed a low frequency of L74M in INSTI-naïve patients [64,68,80]. M50I can be found in 10–25% of INSTI-naïve patients [81]. M50I alone does not negatively impact integrase strand-transfer activity and HIV replication capacity, but in combination with R263K, it increased resistance to DTG by 15.6-fold [81].
The other nonpolymorphic and polymorphic accessory mutations detected were G163R and T97A, which can contribute to a high-level resistance when occurring with Y143 and N155H major INSTI-resistance mutations [30].
In this study, we characterized the distribution of amino acid variants among the 453 HIV-1 subtype C IN sequences from INSTI-naïve individuals. Our results revealed that 64.9% of HIV-1 IN amino acid positions were conserved (<1.0% variability). The conserved position in the NTD, CCD, and CTD were 60%, 66.0%, and 65.8%, respectively. This was comparable to the study by Rhee et al. (2008) that showed 70% (202/288) of IN amino acid positions of the 1500 sequences obtained from INSTI-naïve (ART-naive or ART-experienced) individuals with different subtypes (<1.0% variability) [5]. Similarly, Hackett et al. (2008) also showed that 65% (187/288) of amino acid positions were conserved after analyzing 1304 HIV-1 sequences from groups M, N, and O IN sequences [82].
In general, our results showed that the majority of amino acids involved in key functions of the enzyme, including the zinc-binding HHCC motif, the multimerization of IN subunits, and the binding with the human cellular factor LEDGF/p75 in the catalytic core domain, the catalytic triad DDE [6,22] was highly conserved. The high conservation might have been due to the absence of INSTI pressure. All of our study participants were INSTI-naïve, and INSTI was not used in Ethiopia during our sample collection. However, a highly polymorphic residue in the NTD, CCD, and CTD regions, which might have affected the IN-protein function and interfered with the INSTI binding, were also observed [22,30]. Further long-term treatment follow-up studies are needed to assess the potential impact of NOPs on the evolution of INSTI resistance and viral fitness under the pressure of INSTIs.
It was also interesting to note that 20.5% (93/453) of our study participants were found to harbor a major HIVDR mutation (transmitted and acquired HIVDR) for NRTI, NNRTI, and/or PI in their corresponding PR/RT gene. However, DRM directed toward sites other than IN did not have a significant effect on INSTI susceptibility. In line with our findings, different studies have shown that previous NRTIs mutations appeared to have no impact on the risk of virological failure in patients switched to DTG with NNRTIs [83,84,85,86]. However, this was in contrast to other studies that showed previous exposure to NNRTI, PI, and/or NNRTI induced mutations or increase polymorphisms in the IN gene, highlighting the functional cooperation between viral IN and RT, and/or a potential coevolution of some of their mutations [9,87]. For instance, a study by Ceccherini et al. (2009, 2010) showed a higher frequency of I84V, M154I, and V165I among ART-treated subtype B patients compared to ART-naïve patients, implying that nonsuppressive ART treatment based on other antiretroviral drug classes (NRTI and/or NNRTI) might induce IN polymorphisms [6,9].
However, in our study, no significant difference was found in I84V and M154I prevalence between the ART-naïve and ART-experienced patients (22.6% and 0.5% of I84V and M154I among ART-naïve, and 1.15% and 12.6% among ART-experienced patients, respectively (p = 0.5 and p = 0.4)), while an increased prevalence of V165I was observed among ART-experienced groups (5.45% of V165I and12.64% between the ART-naïve and ART-experienced groups, respectively (p = 0.03)). Furthermore, our comparison of the HIVDR and no-HIVDR groups showed no differences (17.4%, 0.4%, and 7.3% of I84V, M154I, and V165I for the no-HIVDR group; and 22.6%, 1.1%, and 6.5% for the HIVDR group; p = 0.4, p = 0.5, and p = 1, respectively).
The observed differences between this and previous studies might be due to the number of sequences, range of major/minor mutations, and subtypes included in the analysis. However, the lack of a major INSTI mutation among sequences with multiple mutations in the PR/RT gene and the high conservation of amino acids involved in key functions of the IN enzyme did not support the impact of previous ART treatment on INSTI susceptibility.
Our docking analysis further supported our results, and showed no differences between the HIVDR and no-HIVDR groups. In both groups, DTG was successfully docked at a similar position to the original DTG ligand with the best docking score of −6.5 kcal/mol.
The genetic barrier, which is a crucial factor in the development of drug resistance, is defined by a cumulative number of resistance-associated mutations (RAMs) required for the virus to escape drug-selective pressure [53]. It is an important factor that contributes to the development of drug resistance. The variability at the nucleotide level in the IN among the different subtypes could influence the genetic barrier of INSTI drugs. In this study, we explored how the variability between subtypes C and B could affect DTG resistance.
Overall, our analysis of the codon distribution of the selected amino acid position of HIV-1 subtype C and subtype B revealed a similar genetic barrier for the development of DTG resistance between subtype C and B, except at codon position 140, where subtype C had a higher genetic barrier to develop the G140C and G140S mutations compared to subtype B, highlighting a higher genetic barrier for the Q148H/R/K resistance pathway in subtype C. The G140S mutation has been shown to rescue the catalytic defect due to the Q148H mutation, enabling the recovery of viral fitness [88]. A similar high genetic barrier to acquire mutations G140S or G140C has also been described in CRF02_AG compared with subtype B [53,89].
This study was comprehensive, and included both treatment-naïve and treatment-experienced (first- and second-line regimens) patients, and will be a benchmark for INSTI DRM monitoring in Ethiopia. However, our analysis was based on the Sanger dideoxy sequencing method, which does not detect drug-resistance minority variants below 20% of the virus population, and might have underestimated the prevalence of INSTI DRMs among our study participants [90].

5. Conclusions

Our results showed no major clinically relevant INSTI-associated mutations among INSTI-naïve patients regardless of exposure to other antiretroviral agents, supporting the implementation of the wide scale-up of DTG-based regimes in Ethiopia. However, the detection of polymorphisms contributing to INSTI resistance and the expected increased use of DTG-based regimens in Ethiopia warrant the need for continuous surveillance of INSTI resistance. The genetic barrier analysis showed that subtype C had a high genetic barrier to acquiring the G140C and G140S mutations, highlighting that the Q148H/K/R mutation DTG resistance pathway was selected less in subtype C. Moreover, the docking analysis of the dolutegravir showed that protease- and reverse-transcriptase-associated HIVDRMs did not affect the native structure of the HIV-1 integrase.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/v14040729/s1, Table S1: Type of HIV-drug-resistance mutations detected in the HIVDR group (patients with one or more major HIVDR mutation to NRTI, NNRTI, and/or PI) (n = 93), Table S2: Modifications of the 6u8q.pdb HIV-1 integrase structure according to the alignment of both the ADR and PDR sequences.

Author Contributions

Conceptualization, D.A.A., P.B. and P.M.; Data curation, D.A.A.; Formal analysis, D.A.A., S.S. and P.M.; Funding acquisition, F.A, P.B. and P.M.; Investigation, D.A.A., Z.I.S., K.Z., A.G., F.A., S.S. and P.M.; Project administration, D.A.A. and P.M. Supervision, D.A.A. and P.M.; Visualization, Z.I.S.; Writing—original draft, D.A.A.; Writing—review and editing, D.A.A., S.S., P.B. and P.M. All authors revised the manuscript, provided important intellectual content. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the WHO Ethiopia Country Office. The Swedish Research Council (grant numbers 2019-05235 and 2020-02344), the Österlund Foundation, and a donation to P.M. and P.B. through the Medical Faculty, Lund University, also supported the research. The computations were enabled by resources from project SNIC 2021-5-69 provided by the Swedish National Infrastructure for Computing (SNIC) at UPPMAX, partially funded by the Swedish Research Council through grant agreement no. 2018-05973.

Informed Consent Statement

The study was conducted in accordance with the Declaration of Hesinki and approved by Research and Ethical Clearance Committee of EPHI and the National Health Research Ethics Review Committee of the Ministry of Science and Technology of Ethiopia (SERO-60-2016; Version 001, on 11 May 2016). Written informed consent was obtained from all participants for the national HIV-drug-resistance survey.

Data Availability Statement

All the sequences from this study were deposited in the GenBank with accession numbers OM302554–OM303013.

Acknowledgments

We express our deep appreciation to the study participants and all staff involved in the survey. We acknowledge the support given by the WHO and the National Institute of Respiratory Diseases-Mexico (INER) laboratory for HIVDR genotyping. We also acknowledge Mauno Vihinen (Protein Bioinformatics. Lund University) for his assistance with the Chimera software.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. WHO. Guidelines on the Public Health Response to Pretreatment HIV Drug Resistance; WHO: Geneva, Switzerland, 2017; Available online: http://apps.who.int/iris/bitstream/10665/255880/1/9789241550055-eng.pdf (accessed on 20 December 2021).
  2. WHO. Policy Brief: Update of Recommendations on First- and Second-Line Antiretroviral Regimens. 2019. Available online: https://apps.who.int/iris/handle/10665/325892 (accessed on 20 December 2021).
  3. WHO. Dolutegravir (DTG) and the fixed Dose Combination (FDC) of Tenofovir/Lamivudine/Dolutegravir (TLD). 2018. Available online: https://www.who.int/hiv/pub/arv/DTG-TLD-arv_briefing_2018.pdf (accessed on 20 December 2021).
  4. The Lancet HIV. End resistance to dolutegravir roll-out. Lancet HIV 2020, 7, e593. [Google Scholar] [CrossRef]
  5. Rhee, S.Y.; Liu, T.F.; Kiuchi, M.; Zioni, R.; Gifford, R.J.; Holmes, S.P.; Shafer, R.W. Natural variation of HIV-1 group M integrase: Implications for a new class of antiretroviral inhibitors. Retrovirology 2008, 5, 74. [Google Scholar] [CrossRef] [Green Version]
  6. Ceccherini-Silberstein, F.; Malet, I.; D’Arrigo, R.; Antinori, A.; Marcelin, A.G.; Perno, C.F. Characterization and structural analysis of HIV-1 integrase conservation. AIDS Rev. 2009, 11, 17–29. [Google Scholar]
  7. Craigie, R. HIV integrase, a brief overview from chemistry to therapeutics. J. Biol. Chem. 2001, 276, 23213–23216. [Google Scholar] [CrossRef] [Green Version]
  8. Zheng, R.; Jenkins, T.M.; Craigie, R. Zinc folds the N-terminal domain of HIV-1 integrase, promotes multimerization, and enhances catalytic activity. Proc. Natl. Acad. Sci. USA 1996, 93, 13659–13664. [Google Scholar] [CrossRef] [Green Version]
  9. Ceccherini-Silberstein, F.; Malet, I.; Fabeni, L.; Dimonte, S.; Svicher, V.; D’Arrigo, R.; Artese, A.; Costa, G.; Bono, S.; Alcaro, S.; et al. Specific HIV-1 integrase polymorphisms change their prevalence in untreated versus antiretroviral-treated HIV-1-infected patients, all naive to integrase inhibitors. J. Antimicrob. Chemother. 2010, 65, 2305–2318. [Google Scholar] [CrossRef]
  10. Brown, P.O.; Bowerman, B.; Varmus, H.E.; Bishop, J.M. Retroviral integration: Structure of the initial covalent product and its precursor, and a role for the viral IN protein. Proc. Natl. Acad. Sci. USA 1989, 86, 2525–2529. [Google Scholar] [CrossRef] [Green Version]
  11. FDA. FDA-Approved HIV Medicines. 2021. Available online: https://hivinfo.nih.gov/understanding-hiv/fact-sheets/fda-approved-hiv-medicines (accessed on 22 December 2021).
  12. Yang, L.L.; Li, Q.; Zhou, L.B.; Chen, S.Q. Meta-analysis and systematic review of the efficacy and resistance for human immunodeficiency virus type 1 integrase strand transfer inhibitors. Int. J. Antimicrob. Agents 2019, 54, 547–555. [Google Scholar] [CrossRef]
  13. Han, Y.S.; Mesplède, T.; Wainberg, M.A. Differences among HIV-1 subtypes in drug resistance against integrase inhibitors. Infect. Genet. Evol. J. Mol. Epidemiol. Evol. Genet. Infect. Dis. 2016, 46, 286–291. [Google Scholar] [CrossRef]
  14. Dow, D.E.; Bartlett, J.A. Dolutegravir, the Second-Generation of Integrase Strand Transfer Inhibitors (INSTIs) for the Treatment of HIV. Infect Dis. 2014, 3, 83–102. [Google Scholar] [CrossRef] [Green Version]
  15. Lepik, K.J.; Harrigan, P.R.; Yip, B.; Wang, L.; Robbins, M.A.; Zhang, W.W.; Toy, J.; Akagi, L.; Lima, V.D.; Guillemi, S.; et al. Emergent drug resistance with integrase strand transfer inhibitor-based regimens. AIDS 2017, 31, 1425–1434. [Google Scholar] [CrossRef]
  16. McGee, K.S.; Okeke, N.L.; Hurt, C.B.; McKellar, M.S. Canary in the Coal Mine? Transmitted Mutations Conferring Resistance to All Integrase Strand Transfer Inhibitors in a Treatment-Naive Patient. Open Forum Infect. Dis. 2018, 5, ofy294. [Google Scholar] [CrossRef]
  17. Oliveira, M.; Ibanescu, R.I.; Anstett, K.; Mésplède, T.; Routy, J.P.; Robbins, M.A.; Brenner, B.G. Selective resistance profiles emerging in patient-derived clinical isolates with cabotegravir, bictegravir, dolutegravir, and elvitegravir. Retrovirology 2018, 15, 56. [Google Scholar] [CrossRef] [Green Version]
  18. Young, B.; Fransen, S.; Greenberg, K.S.; Thomas, A.; Martens, S.; St Clair, M.; Petropoulos, C.J.; Ha, B. Transmission of integrase strand-transfer inhibitor multidrug-resistant HIV-1: Case report and response to raltegravir-containing antiretroviral therapy. Antivir 2011, 16, 253–256. [Google Scholar] [CrossRef] [Green Version]
  19. Seatla, K.K.; Maruapula, D.; Choga, W.T.; Ntsipe, T.; Mathiba, N.; Mogwele, M.; Kapanda, M.; Nkomo, B.; Ramaabya, D.; Makhema, J.; et al. HIV-1 Subtype C Drug Resistance Mutations in Heavily Treated Patients Failing Integrase Strand Transfer Inhibitor-Based Regimens in Botswana. Viruses 2021, 13, 594. [Google Scholar] [CrossRef]
  20. Bailey, A.J.; Rhee, S.Y.; Shafer, R.W. Integrase Strand Transfer Inhibitor Resistance in Integrase Strand Transfer Inhibitor-Naive Persons. AIDS Res. Hum. Retrovir. 2021, 37, 736–743. [Google Scholar] [CrossRef]
  21. Casadellà, M.; Santos, J.R.; Noguera-Julian, M.; Micán-Rivera, R.; Domingo, P.; Antela, A.; Portilla, J.; Sanz, J.; Montero-Alonso, M.; Navarro, J.; et al. Primary resistance to integrase strand transfer inhibitors in Spain using ultrasensitive HIV-1 genotyping. J. Antimicrob. Chemother. 2020, 75, 3517–3524. [Google Scholar] [CrossRef]
  22. Semengue, E.N.J.; Armenia, D.; Inzaule, S.; Santoro, M.M.; Dambaya, B.; Takou, D.; Teto, G.; Nka, A.D.; Yagai, B.; Fabeni, L.; et al. Baseline integrase drug resistance mutations and conserved regions across HIV-1 clades in Cameroon: Implications for transition to dolutegravir in resource-limited settings. J. Antimicrob. Chemother. 2021, 76, 1277–1285. [Google Scholar] [CrossRef]
  23. Lübke, N.; Jensen, B.; Hüttig, F.; Feldt, T.; Walker, A.; Thielen, A.; Däumer, M.; Obermeier, M.; Kaiser, R.; Knops, E.; et al. Failure of Dolutegravir First-Line ART with Selection of Virus Carrying R263K and G118R. N. Engl. J. Med. 2019, 381, 887–889. [Google Scholar] [CrossRef]
  24. Ndashimye, E.; Avino, M.; Olabode, A.S.; Poon, A.F.Y.; Gibson, R.M.; Li, Y.; Meadows, A.; Tan, C.; Reyes, P.S.; Kityo, C.M.; et al. Accumulation of integrase strand transfer inhibitor resistance mutations confers high-level resistance to dolutegravir in non-B subtype HIV-1 strains from patients failing raltegravir in Uganda. J. Antimicrob. Chemother. 2020, 75, 3525–3533. [Google Scholar] [CrossRef]
  25. Pena, M.J.; Chueca, N.; D’Avolio, A.; Zarzalejos, J.M.; Garcia, F. Virological Failure in HIV to Triple Therapy With Dolutegravir-Based Firstline Treatment: Rare but Possible. Open Forum Infect. Dis. 2019, 6, ofy332. [Google Scholar] [CrossRef] [Green Version]
  26. Blanco, J.L.; Marcelin, A.G.; Katlama, C.; Martinez, E. Dolutegravir resistance mutations: Lessons from monotherapy studies. Curr. Opin Infect. Dis. 2018, 31, 237–245. [Google Scholar] [CrossRef]
  27. Brenner, B.G.; Thomas, R.; Blanco, J.L.; Ibanescu, R.I.; Oliveira, M.; Mesplède, T.; Golubkov, O.; Roger, M.; Garcia, F.; Martinez, E.; et al. Development of a G118R mutation in HIV-1 integrase following a switch to dolutegravir monotherapy leading to cross-resistance to integrase inhibitors. J. Antimicrob. Chemother. 2016, 71, 1948–1953. [Google Scholar] [CrossRef]
  28. Fulcher, J.A.; Du, Y.; Zhang, T.H.; Sun, R.; Landovitz, R.J. Emergence of Integrase Resistance Mutations During Initial Therapy Containing Dolutegravir. Clin. Infect. Dis. 2018, 67, 791–794. [Google Scholar] [CrossRef]
  29. Theys, K.; Libin, P.J.K.; Van Laethem, K.; Abecasis, A.B. An Evolutionary Model-Based Approach To Quantify the Genetic Barrier to Drug Resistance in Fast-Evolving Viruses and Its Application to HIV-1 Subtypes and Integrase Inhibitors. Antimicrob. Agents Chemother. 2019, 63, e00539-19. [Google Scholar] [CrossRef] [Green Version]
  30. Mikasi, S.G.; Isaacs, D.; Chitongo, R.; Ikomey, G.M.; Jacobs, G.B.; Cloete, R. Interaction analysis of statistically enriched mutations identified in Cameroon recombinant subtype CRF02_AG that can influence the development of Dolutegravir drug resistance mutations. BMC Infect. Dis. 2021, 21, 379. [Google Scholar] [CrossRef]
  31. Doyle, T.; Dunn, D.T.; Ceccherini-Silberstein, F.; De Mendoza, C.; Garcia, F.; Smit, E.; Fearnhill, E.; Marcelin, A.G.; Martinez-Picado, J.; Kaiser, R.; et al. Integrase inhibitor (INI) genotypic resistance in treatment-naive and raltegravir-experienced patients infected with diverse HIV-1 clades. J. Antimicrob. Chemother. 2015, 70, 3080–3086. [Google Scholar] [CrossRef] [Green Version]
  32. Rogers, L.; Obasa, A.E.; Jacobs, G.B.; Sarafianos, S.G.; Sönnerborg, A.; Neogi, U.; Singh, K. Structural Implications of Genotypic Variations in HIV-1 Integrase From Diverse Subtypes. Front. Microbiol. 2018, 9, 1754. [Google Scholar] [CrossRef] [Green Version]
  33. Bar-Magen, T.; Donahue, D.A.; McDonough, E.I.; Kuhl, B.D.; Faltenbacher, V.H.; Xu, H.; Michaud, V.; Sloan, R.D.; Wainberg, M.A. HIV-1 subtype B and C integrase enzymes exhibit differential patterns of resistance to integrase inhibitors in biochemical assays. Aids 2010, 24, 2171–2179. [Google Scholar] [CrossRef]
  34. Kirichenko, A.; Lapovok, I.; Baryshev, P.; van de Vijver, D.; van Kampen, J.J.A.; Boucher, C.A.B.; Paraskevis, D.; Kireev, D. Genetic Features of HIV-1 Integrase Sub-Subtype A6 Predominant in Russia and Predicted Susceptibility to INSTIs. Viruses 2020, 12, 838. [Google Scholar] [CrossRef]
  35. Quashie, P.K.; Mesplède, T.; Han, Y.S.; Oliveira, M.; Singhroy, D.N.; Fujiwara, T.; Underwood, M.R.; Wainberg, M.A. Characterization of the R263K mutation in HIV-1 integrase that confers low-level resistance to the second-generation integrase strand transfer inhibitor dolutegravir. J. Virol. 2012, 86, 2696–2705. [Google Scholar] [CrossRef] [Green Version]
  36. Rhee, S.Y.; Grant, P.M.; Tzou, P.L.; Barrow, G.; Harrigan, P.R.; Ioannidis, J.P.A.; Shafer, R.W. A systematic review of the genetic mechanisms of dolutegravir resistance. J. Antimicrob. Chemother. 2019, 74, 3135–3149. [Google Scholar] [CrossRef]
  37. Mouscadet, J.F.; Delelis, O.; Marcelin, A.G.; Tchertanov, L. Resistance to HIV-1 integrase inhibitors: A structural perspective. Drug Resist. Updates 2010, 13, 139–150. [Google Scholar] [CrossRef]
  38. EPHI. HIV Related Estimats and Projections in Ethiopia for the Year-2019. 2020. Available online: http://repository.iifphc.org/bitstream/handle/123456789/1069/HIV_estimation_and_projection_for_Ethiopia_2019.pdf?sequence=1&isAllowed=y (accessed on 25 December 2021).
  39. Arimide, D.A.; Esquivel-Gómez, L.R.; Kebede, Y.; Sasinovich, S.; Balcha, T.; Björkman, P.; Kühnert, D.; Medstrand, P. Molecular Epidemiology and Transmission Dynamics of the HIV-1 Epidemic in Ethiopia: Epidemic Decline Coincided With Behavioral Interventions Before ART Scale-Up. Front. Microbiol. 2022, 13, 821006. [Google Scholar] [CrossRef]
  40. MOH. National Consolidated Guidelines for Comprehensive HIV Prevention, Care and Treatment. 2018. Available online: https://www.afro.who.int/sites/default/files/2019-04/National%20Comprehensive%20HIV%20Care%20%20Guideline%202018.pdf (accessed on 15 December 2021).
  41. WHO. HIV Drug Resistance Surveillance Guidance: 2015 Update. 2015. Available online: https://www.who.int/publications/i/item/978-92-4-151009-7 (accessed on 27 December 2021).
  42. WHO. World Health Organization Global Strategy for the Surveillance and Monitoring of HIV Drug Resistance; WHO Press: Geneva, Switzerland, 2012. [Google Scholar]
  43. Van Laethem, K.; Schrooten, Y.; Covens, K.; Dekeersmaeker, N.; De Munter, P.; Van Wijngaerden, E.; Van Ranst, M.; Vandamme, A.M. A genotypic assay for the amplification and sequencing of integrase from diverse HIV-1 group M subtypes. J. Virol. Methods 2008, 153, 176–181. [Google Scholar] [CrossRef]
  44. Woods, C.K.; Brumme, C.J.; Liu, T.F.; Chui, C.K.; Chu, A.L.; Wynhoven, B.; Hall, T.A.; Trevino, C.; Shafer, R.W.; Harrigan, P.R. Automating HIV drug resistance genotyping with RECall, a freely accessible sequence analysis tool. J. Clin. Microbiol. 2012, 50, 1936–1942. [Google Scholar] [CrossRef] [Green Version]
  45. Pineda-Peña, A.C.; Faria, N.R.; Imbrechts, S.; Libin, P.; Abecasis, A.B.; Deforche, K.; Gómez-López, A.; Camacho, R.J.; de Oliveira, T.; Vandamme, A.M. Automated subtyping of HIV-1 genetic sequences for clinical and surveillance purposes: Performance evaluation of the new REGA version 3 and seven other tools. Infect. Genet. Evol. J. Mol. Epidemiol. Evol. Genet. Infect. Dis. 2013, 19, 337–348. [Google Scholar] [CrossRef] [Green Version]
  46. Struck, D.; Lawyer, G.; Ternes, A.M.; Schmit, J.C.; Bercoff, D.P. COMET: Adaptive context-based modeling for ultrafast HIV-1 subtype identification. Nucleic. Acids Res. 2014, 42, e144. [Google Scholar] [CrossRef]
  47. Martin, D.P.; Lemey, P.; Lott, M.; Moulton, V.; Posada, D.; Lefeuvre, P. RDP3: A flexible and fast computer program for analyzing recombination. Bioinformatics 2010, 26, 2462–2463. [Google Scholar] [CrossRef]
  48. Katoh, K.; Kuma, K.; Miyata, T.; Toh, H. Improvement in the accuracy of multiple sequence alignment program MAFFT. Genome Inf. 2005, 16, 22–33. [Google Scholar]
  49. Hall, T.A. BioEdit: A user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symp. Ser. 1999, 41, 95–98. [Google Scholar]
  50. Tippmann, H.F. Analysis for free: Comparing programs for sequence analysis. Brief. Bioinform. 2004, 5, 82–87. [Google Scholar] [CrossRef]
  51. Guindon, S.; Dufayard, J.F.; Lefort, V.; Anisimova, M.; Hordijk, W.; Gascuel, O. New algorithms and methods to estimate maximum-likelihood phylogenies: Assessing the performance of PhyML 3.0. Syst. Biol. 2010, 59, 307–321. [Google Scholar] [CrossRef] [Green Version]
  52. Guindon, S.; Lethiec, F.; Duroux, P.; Gascuel, O. PHYML Online—A web server for fast maximum likelihood-based phylogenetic inference. Nucleic Acids Res. 2005, 33, W557-9. [Google Scholar] [CrossRef] [Green Version]
  53. Nguyen, H.L.; Ruxrungtham, K.; Delaugerre, C. Genetic barrier to the development of resistance to integrase inhibitors in HIV-1 subtypes CRF01_AE and B. Intervirology 2012, 55, 287–295. [Google Scholar] [CrossRef]
  54. Li, M.; Chen, X.; Wang, H.; Jurado, K.A.; Engelman, A.N.; Craigie, R. A Peptide Derived from Lens Epithelium-Derived Growth Factor Stimulates HIV-1 DNA Integration and Facilitates Intasome Structural Studies. J. Mol. Biol. 2020, 432, 2055–2066. [Google Scholar] [CrossRef]
  55. Trott, O.; Olson, A.J. AutoDock Vina: Improving the speed and accuracy of docking with a new scoring function, efficient optimization, and multithreading. J. Comput. Chem. 2010, 31, 455–461. [Google Scholar] [CrossRef] [Green Version]
  56. Eberhardt, J.; Santos-Martins, D.; Tillack, A.F.; Forli, S. AutoDock Vina 1.2.0: New Docking Methods, Expanded Force Field, and Python Bindings. J. Chem. Inf. Model. 2021, 61, 3891–3898. [Google Scholar] [CrossRef]
  57. Lu, R.; Limón, A.; Ghory, H.Z.; Engelman, A. Genetic analyses of DNA-binding mutants in the catalytic core domain of human immunodeficiency virus type 1 integrase. J. Virol. 2005, 79, 2493–2505. [Google Scholar] [CrossRef] [Green Version]
  58. Inzaule, S.C.; Hamers, R.L.; Noguera-Julian, M.; Casadellà, M.; Parera, M.; Rinke de Wit, T.F.; Paredes, R. Primary resistance to integrase strand transfer inhibitors in patients infected with diverse HIV-1 subtypes in sub-Saharan Africa. J. Antimicrob. Chemother. 2018, 73, 1167–1172. [Google Scholar] [CrossRef] [Green Version]
  59. Obasa, A.E.; Mikasi, S.G.; Brado, D.; Cloete, R.; Singh, K.; Neogi, U.; Jacobs, G.B. Drug Resistance Mutations Against Protease, Reverse Transcriptase and Integrase Inhibitors in People Living With HIV-1 Receiving Boosted Protease Inhibitors in South Africa. Front. Microbiol. 2020, 11, 438. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Fish, M.Q.; Hewer, R.; Wallis, C.L.; Venter, W.D.; Stevens, W.S.; Papathanasopoulos, M.A. Natural polymorphisms of integrase among HIV type 1-infected South African patients. AIDS Res. Hum. Retrovir. 2010, 26, 489–493. [Google Scholar] [CrossRef] [PubMed]
  61. Oliveira, M.F.; Ramalho, D.B.; Abreu, C.M.; Vubil, A.; Mabunda, N.; Ismael, N.; Francisco, C.; Jani, I.V.; Tanuri, A. Genetic diversity and naturally polymorphisms in HIV type 1 integrase isolates from Maputo, Mozambique: Implications for integrase inhibitors. AIDS Res. Hum. Retrovir. 2012, 28, 1788–1792. [Google Scholar] [CrossRef] [Green Version]
  62. Mulu, A.; Maier, M.; Liebert, U.G. Lack of integrase inhibitors associated resistance mutations among HIV-1C isolates. J. Transl. Med. 2015, 13, 377. [Google Scholar] [CrossRef] [Green Version]
  63. Rangel, H.R.; Garzaro, D.; Fabbro, R.; Martinez, N.; Ossenkop, J.; Torres, J.R.; Gutiérrez, C.R.; Pujol, F.H. Absence of primary integrase resistance mutations in HIV type 1-infected patients in Venezuela. AIDS Res. Hum. Retrovir. 2010, 26, 923–926. [Google Scholar] [CrossRef]
  64. Brado, D.; Obasa, A.E.; Ikomey, G.M.; Cloete, R.; Singh, K.; Engelbrecht, S.; Neogi, U.; Jacobs, G.B. Analyses of HIV-1 integrase sequences prior to South African national HIV-treatment program and available of integrase inhibitors in Cape Town, South Africa. Sci. Rep. 2018, 8, 4709. [Google Scholar] [CrossRef] [PubMed]
  65. Kim, J.Y.; Kim, E.J.; Choi, J.Y.; Kwon, O.K.; Kim, G.J.; Choi, S.Y.; Kim, S.S. Genetic variation of the HIV-1 integrase region in newly diagnosed anti-retroviral drug-naïve patients with HIV/AIDS in Korea. Clin. Microbiol. Infect. 2011, 17, 1155–1159. [Google Scholar] [CrossRef] [Green Version]
  66. Kotaki, T.; Khairunisa, S.Q.; Sukartiningrum, S.D.; Witaningrum, A.M.; Rusli, M.; Diansyah, M.N.; Arfijanto, M.V.; Rahayu, R.P.; Nasronudin; Kameoka, M. Detection of drug resistance-associated mutations in human immunodeficiency virus type 1 integrase derived from drug-naive individuals in Surabaya, Indonesia. AIDS Res. Hum. Retrovir. 2014, 30, 489–492. [Google Scholar] [CrossRef]
  67. Arruda, L.B.; Fonseca, L.A.; Duarte, A.J.; Casseb, J. Genetic diversity on the integrase region of the pol gene among HIV type 1-infected patients naive for integrase inhibitors in São Paulo City, Brazil. AIDS Res. Hum. Retrovir. 2010, 26, 105–107. [Google Scholar] [CrossRef] [Green Version]
  68. Casadellà, M.; van Ham, P.M.; Noguera-Julian, M.; van Kessel, A.; Pou, C.; Hofstra, L.M.; Santos, J.R.; Garcia, F.; Struck, D.; Alexiev, I.; et al. Primary resistance to integrase strand-transfer inhibitors in Europe. J. Antimicrob. Chemother. 2015, 70, 2885–2888. [Google Scholar]
  69. Meixenberger, K.; Yousef, K.P.; Smith, M.R.; Somogyi, S.; Fiedler, S.; Bartmeyer, B.; Hamouda, O.; Bannert, N.; von Kleist, M.; Kücherer, C. Molecular evolution of HIV-1 integrase during the 20 years prior to the first approval of integrase inhibitors. Virol. J. 2017, 14, 223. [Google Scholar] [CrossRef] [Green Version]
  70. Tostevin, A.; White, E.; Dunn, D.; Croxford, S.; Delpech, V.; Williams, I.; Asboe, D.; Pozniak, A.; Churchill, D.; Geretti, A.M.; et al. Recent trends and patterns in HIV-1 transmitted drug resistance in the United Kingdom. HIV Med. 2017, 18, 204–213. [Google Scholar] [CrossRef] [PubMed]
  71. Inzaule, S.C.; Hamers, R.L.; Noguera-Julian, M.; Casadellà, M.; Parera, M.; Kityo, C.; Steegen, K.; Naniche, D.; Clotet, B.; Rinke de Wit, T.F.; et al. Clinically relevant thresholds for ultrasensitive HIV drug resistance testing: A multi-country nested case-control study. Lancet HIV 2018, 5, e638–e646. [Google Scholar] [CrossRef]
  72. Ndashimye, E.; Avino, M.; Kyeyune, F.; Nankya, I.; Gibson, R.M.; Nabulime, E.; Poon, A.F.Y.; Kityo, C.; Mugyenyi, P.; Quiñones-Mateu, M.E.; et al. Absence of HIV-1 Drug Resistance Mutations Supports the Use of Dolutegravir in Uganda. AIDS Res. Hum. Retrovir. 2018, 34, 404–414. [Google Scholar] [CrossRef] [PubMed]
  73. Karade, S.; Sen, S.; Sashindran, V.K. Absence of Integrase Strand Transfer Inhibitor Associated Resistance in Antiretroviral Therapy Naïve and Experienced Individuals from Western India. AIDS Res. Hum. Retrovir. 2019, 35, 567–571. [Google Scholar] [CrossRef] [Green Version]
  74. Kityo, C.; Thompson, J.; Nankya, I.; Hoppe, A.; Ndashimye, E.; Warambwa, C.; Mambule, I.; van Oosterhout, J.J.; Wools-Kaloustian, K.; Bertagnolio, S.; et al. HIV Drug Resistance Mutations in Non-B Subtypes After Prolonged Virological Failure on NNRTI-Based First-Line Regimens in Sub-Saharan Africa. J. Acquir. Immune Defic. Syndr. 2017, 75, e45–e54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Lan, Y.; Li, L.; Chen, W.; Deng, X.; Li, J.; Fan, Q.; Cai, X.; Cai, W.; Hu, F. Absence of Integrase Inhibitor-Associated Resistance Among Antiretroviral Therapy-Naïve HIV-1-Infected Adults in Guangdong Province, China, in 2018. Infect. Drug Resist. 2020, 13, 4389–4394. [Google Scholar] [CrossRef]
  76. Anstett, K.; Cutillas, V.; Fusco, R.; Mesplède, T.; Wainberg, M.A. Polymorphic substitution E157Q in HIV-1 integrase increases R263K-mediated dolutegravir resistance and decreases DNA binding activity. J. Antimicrob. Chemother. 2016, 71, 2083–2088. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Fun, A.; Van Baelen, K.; van Lelyveld, S.F.; Schipper, P.J.; Stuyver, L.J.; Wensing, A.M.; Nijhuis, M. Mutation Q95K enhances N155H-mediated integrase inhibitor resistance and improves viral replication capacity. J. Antimicrob. Chemother. 2010, 65, 2300–2304. [Google Scholar] [CrossRef] [Green Version]
  78. Garrido, C.; Villacian, J.; Zahonero, N.; Pattery, T.; Garcia, F.; Gutierrez, F.; Caballero, E.; Van Houtte, M.; Soriano, V.; de Mendoza, C. Broad phenotypic cross-resistance to elvitegravir in HIV-infected patients failing on raltegravir-containing regimens. Antimicrob. Agents Chemother. 2012, 56, 2873–2878. [Google Scholar] [CrossRef] [Green Version]
  79. Temesgen, Z.; Siraj, D.S. Raltegravir: First in class HIV integrase inhibitor. Clin. Risk Manag. 2008, 4, 493–500. [Google Scholar] [CrossRef] [Green Version]
  80. Passaes, C.B.; Guimarães, M.L.; Fernandez, S.L.; Lorete Rdos, S.; Teixeira, S.L.; Fernandez, J.C.; Morgado, M.G. Lack of primary mutations associated with integrase inhibitors among HIV-1 subtypes B, C, and F circulating in Brazil. J. Acquir. Immune Defic. Syndr. 2009, 51, 7–12. [Google Scholar] [CrossRef] [PubMed]
  81. Wares, M.; Mesplède, T.; Quashie, P.K.; Osman, N.; Han, Y.; Wainberg, M.A. The M50I polymorphic substitution in association with the R263K mutation in HIV-1 subtype B integrase increases drug resistance but does not restore viral replicative fitness. Retrovirology 2014, 11, 7. [Google Scholar] [CrossRef] [Green Version]
  82. Hackett, J.; Harris, B.; Holzmayer, V.; Yamaguchi, J.; Luk, K.C.; Brennan, C.; Schochetman, G.; Devare, S.; Swanson, P. Naturally occurring polymorphisms in HIV-1 Group M, N, and O Integrase: Implications for integrase inhibitors. In Proceedings of the Fifteenth Conference on Retroviruses and Opportunistic Infections, Boston, MA, USA, 2–6 February 2008. [Google Scholar]
  83. Giacomelli, A.; Lai, A.; Franzetti, M.; Maggiolo, F.; Di Giambenedetto, S.; Borghi, V.; Francisci, D.; Magnani, G.; Pecorari, M.; Monno, L.; et al. No impact of previous NRTIs resistance in HIV positive patients switched to DTG+2NRTIs under virological control: Time of viral suppression makes the difference. Antivir. Res. 2019, 172, 104635. [Google Scholar] [CrossRef] [PubMed]
  84. Olearo, F.; Nguyen, H.; Bonnet, F.; Yerly, S.; Wandeler, G.; Stoeckle, M.; Cavassini, M.; Scherrer, A.; Costagiola, D.; Schmid, P.; et al. Impact of the M184V/I Mutation on the Efficacy of Abacavir/Lamivudine/Dolutegravir Therapy in HIV Treatment-Experienced Patients. Open Forum Infect. Dis. 2019, 6, ofz330. [Google Scholar] [CrossRef]
  85. Andreatta, K.; Willkom, M.; Martin, R.; Chang, S.; Wei, L.; Liu, H.; Liu, Y.-P.; Graham, H.; Quirk, E.; Martin, H.; et al. Switching to bictegravir/emtricitabine/tenofovir alafenamide maintained HIV-1 RNA suppression in participants with archived antiretroviral resistance including M184V/I. J. Antimicrob. Chemother. 2019, 74, 3555–3564. [Google Scholar] [CrossRef] [Green Version]
  86. Chen, G.J.; Sun, H.Y.; Chang, S.Y.; Cheng, A.; Huang, Y.S.; Lin, K.Y.; Huang, Y.C.; Su, Y.C.; Liu, W.C.; Hung, C.C. Effectiveness of switching from protease inhibitors to dolutegravir in combination with nucleoside reverse transcriptase inhibitors as maintenance antiretroviral therapy among HIV-positive patients. Int. J. Antimicrob. Agents 2019, 54, 35–42. [Google Scholar] [CrossRef]
  87. Van Hal, S.J.; Herring, B.; Deris, Z.; Wang, B.; Saksena, N.K.; Dwyer, D.E. HIV-1 integrase polymorphisms are associated with prior antiretroviral drug exposure. Retrovirology 2009, 6, 12. [Google Scholar] [CrossRef] [Green Version]
  88. Brenner, B.G.; Lowe, M.; Moisi, D.; Hardy, I.; Gagnon, S.; Charest, H.; Baril, J.G.; Wainberg, M.A.; Roger, M. Subtype diversity associated with the development of HIV-1 resistance to integrase inhibitors. J. Med. Virol. 2011, 83, 751–759. [Google Scholar] [CrossRef]
  89. Maïga, A.I.; Malet, I.; Soulie, C.; Derache, A.; Koita, V.; Amellal, B.; Tchertanov, L.; Delelis, O.; Morand-Joubert, L.; Mouscadet, J.F.; et al. Genetic barriers for integrase inhibitor drug resistance in HIV type-1 B and CRF02_AG subtypes. Antivir 2009, 14, 123–129. [Google Scholar]
  90. Telele, N.F.; Kalu, A.W.; Gebre-Selassie, S.; Fekade, D.; Abdurahman, S.; Marrone, G.; Neogi, U.; Tegbaru, B.; Sönnerborg, A. Pretreatment drug resistance in a large countrywide Ethiopian HIV-1C cohort: A comparison of Sanger and high-throughput sequencing. Sci. Rep. 2018, 8, 7556. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Maximum-likelihood phylogenetic tree of HIV-1 viral strains circulating in Ethiopia using integrase (IN) sequences.
Figure 1. Maximum-likelihood phylogenetic tree of HIV-1 viral strains circulating in Ethiopia using integrase (IN) sequences.
Viruses 14 00729 g001
Figure 2. Flow chart of study-participant selection and INSTI drug resistance among INSTI-naïve patients (n = 460). Abbreviations: PR/RT, protease and reverse transcriptase gene; IN, integrase; INSTI, integrase strand-transfer inhibitors; HIVDRM, HIV-drug-resistance mutations.
Figure 2. Flow chart of study-participant selection and INSTI drug resistance among INSTI-naïve patients (n = 460). Abbreviations: PR/RT, protease and reverse transcriptase gene; IN, integrase; INSTI, integrase strand-transfer inhibitors; HIVDRM, HIV-drug-resistance mutations.
Viruses 14 00729 g002
Figure 3. Alignment of Ethiopian HIV-1 subtype C integrase (IN) consensus sequence. The consensus sequence from the ART-naïve sequences (n = 367) is represented as ART_Naive, and that from ART-experienced (n = 87) is represented as ART_Expo. The consensus sequence from the sequence with no HIVDR mutation in the protease/reverse transcriptase (PR/RT) gene (n = 234) is represented as No_HIVDR, while that with one or more major mutation in PR/RT is represented as HIVDR. Positions with more than one amino acid are both represented. HXB2 represents the consensus HIV-1 subtype B reference sequence from the LANL database (accession number: K03455).
Figure 3. Alignment of Ethiopian HIV-1 subtype C integrase (IN) consensus sequence. The consensus sequence from the ART-naïve sequences (n = 367) is represented as ART_Naive, and that from ART-experienced (n = 87) is represented as ART_Expo. The consensus sequence from the sequence with no HIVDR mutation in the protease/reverse transcriptase (PR/RT) gene (n = 234) is represented as No_HIVDR, while that with one or more major mutation in PR/RT is represented as HIVDR. Positions with more than one amino acid are both represented. HXB2 represents the consensus HIV-1 subtype B reference sequence from the LANL database (accession number: K03455).
Viruses 14 00729 g003
Figure 4. Distribution of variants among HIV-1 subtype C integrase sequences. The number at the top is the amino acid position in the integrase gene (1–288), and the consensus subtype B sequence is indicated below the number. Beneath the consensus, the number indicates the number of sequences containing the amino acid at the indicated position. The variant amino acid at each position is indicated along with number of sequences with that amino acid (superscript). The HHCC zinc-binding motifs are indicated by **, the amino acid of the DDE active sites are indicated by red boxes. The N-terminal domain is indicated in gray at positions 1–50, while the catalytic core domain (CCD) at positions 51–212 is indicated in orange, and the C-terminal domain (CTD) at positions 213–288 is indicated in blue.
Figure 4. Distribution of variants among HIV-1 subtype C integrase sequences. The number at the top is the amino acid position in the integrase gene (1–288), and the consensus subtype B sequence is indicated below the number. Beneath the consensus, the number indicates the number of sequences containing the amino acid at the indicated position. The variant amino acid at each position is indicated along with number of sequences with that amino acid (superscript). The HHCC zinc-binding motifs are indicated by **, the amino acid of the DDE active sites are indicated by red boxes. The N-terminal domain is indicated in gray at positions 1–50, while the catalytic core domain (CCD) at positions 51–212 is indicated in orange, and the C-terminal domain (CTD) at positions 213–288 is indicated in blue.
Viruses 14 00729 g004
Figure 5. Prevalence of integrase highly polymorphic mutation in INSTI-naïve patients. The comparison was done using sequences downloaded from the Los Alamos National Library (LANL) database for subtype B (n = 1884), HIV-1 subtype C (n = 1410), and Ethiopian subtype C (n = 453).
Figure 5. Prevalence of integrase highly polymorphic mutation in INSTI-naïve patients. The comparison was done using sequences downloaded from the Los Alamos National Library (LANL) database for subtype B (n = 1884), HIV-1 subtype C (n = 1410), and Ethiopian subtype C (n = 453).
Viruses 14 00729 g005
Figure 6. Alignment of integrase (IN) consensus sequence from LANL database (https://www.hiv.lanl.gov (accessed on 25 July 2021)). The consensus sequence from the global HIV-1 subtype C (n = 1410 sequences) is represented as GLOBAL C, the global HIV-1 subtype B (n = 1884 sequences) is represented as GLOBAL_B, and the Ethiopian subtype C sequence (n = 453 sequences) is represented as ETHIOPIA_C. Positions with more than one amino acid are both represented. HXB2 represents the consensus HIV-1 subtype B reference sequence from the LANL database (accession number: K03455).
Figure 6. Alignment of integrase (IN) consensus sequence from LANL database (https://www.hiv.lanl.gov (accessed on 25 July 2021)). The consensus sequence from the global HIV-1 subtype C (n = 1410 sequences) is represented as GLOBAL C, the global HIV-1 subtype B (n = 1884 sequences) is represented as GLOBAL_B, and the Ethiopian subtype C sequence (n = 453 sequences) is represented as ETHIOPIA_C. Positions with more than one amino acid are both represented. HXB2 represents the consensus HIV-1 subtype B reference sequence from the LANL database (accession number: K03455).
Viruses 14 00729 g006
Figure 7. Visualization of HIV-1 IN amino acid sequences from ART-naive (PDR) and ART-experienced drug-resistant (ADR) individuals. (a) Heat map of amino acid identities of HIV-1 IN amino acid sequences from PDR and ADR individuals. The heat map was generated using a percent identity matrix table created in Clustal 12.1, and the heat map was visualized in GraphPad Prism 8. (b) For in silico predictions, a molecular model of the monomeric HIV-1 integrase structure 6u8q was used and modified based on the multiple-sequence alignment of amino acid sequences of 10 ADR and 10 PDR sequences. Figures were created in Chimera (https://www.cgl.ucsf.edu/chimera https://www.cgl.ucsf.edu/chimera (accessed on 2 November 2021)). PDR (salmon) and ADR (cornflower blue) HIV-1 monomers are represented as ribbons, the catalytic triad is represented with sticks in light grey, and magnesium is represented in orange-red. The PDR structure was moved at 0.01 Å on the x-axis. (c) Surface views of the structure and the validation of docking were conducted using Autodock Vina. Dolutegravir is represented with sticks; blue shows the original coordinates, and purple shows the docking mode of dolutegravir in Autodock Vina.
Figure 7. Visualization of HIV-1 IN amino acid sequences from ART-naive (PDR) and ART-experienced drug-resistant (ADR) individuals. (a) Heat map of amino acid identities of HIV-1 IN amino acid sequences from PDR and ADR individuals. The heat map was generated using a percent identity matrix table created in Clustal 12.1, and the heat map was visualized in GraphPad Prism 8. (b) For in silico predictions, a molecular model of the monomeric HIV-1 integrase structure 6u8q was used and modified based on the multiple-sequence alignment of amino acid sequences of 10 ADR and 10 PDR sequences. Figures were created in Chimera (https://www.cgl.ucsf.edu/chimera https://www.cgl.ucsf.edu/chimera (accessed on 2 November 2021)). PDR (salmon) and ADR (cornflower blue) HIV-1 monomers are represented as ribbons, the catalytic triad is represented with sticks in light grey, and magnesium is represented in orange-red. The PDR structure was moved at 0.01 Å on the x-axis. (c) Surface views of the structure and the validation of docking were conducted using Autodock Vina. Dolutegravir is represented with sticks; blue shows the original coordinates, and purple shows the docking mode of dolutegravir in Autodock Vina.
Viruses 14 00729 g007
Table 1. Prevalence of integrase accessory mutations detected and their ART status.
Table 1. Prevalence of integrase accessory mutations detected and their ART status.
No.Sequence IDART RegimenAgeGenderCD4+ T-Cell Count (Cells/mm3)Viral Load (Copies/mL)INSTI Accessory Mutation
1ETH-0186Naive40M35962,118E157Q
2ETH-0232Naive42M83--G163R
3ETH-0343Naive40M4226,531E157Q
4ETH-0358Naive35M69418,611G149A, E157Q
5ETH-0366Naive27F17562,517E157Q
6ETH-0380Naive34M16397,306E157Q
7ETH-0396Naive25F34147,435G163R
8ETH-0410Naive38F53811,458E157Q
9ETH-0493Naive45M1647130Q95K
10ETH-0508Naive39M150--E157Q
11ETH-0545Naive30F----E157Q
12ETH-0609Naive28F8952002T97A
13ETH-0622Naive21F236155,331T97A
14ETH-0631Naive35F50295,532E157Q
15ETH-0695Naive35F--6465E157Q
16ETH-0750TDF+3TC+EFV46M--18,681Q95K
17ETH-0815TDF+3TC+EFV50M384--Q95K
18ETH-0839ABC+3TC+ATV/r40M4321432G163K
19ETH-0843AZT+3TC+LPV/r39M7334752G140E
20ETH-0879TDF+3TC+ATV/r50F6552667E157Q
Abbreviations: age, in years; F, female, M, male; ART, antiretroviral therapy; INSTI, integrase strand-transfer inhibitor; 3TC, lamiduvine; TDF, tenofovir, AZT, zidovudine; EFV, efavirenz; LPV/r, lopinavir/ritonavir; ATV/r, atazanavir/ritonavir; Naïve, ART-naive; ”--”, missing data. CD4+ T in cells/mm3; HIV RNA in copies/mL.
Table 2. Analysis of genetic barrier based on the minimum number of transitions and transversions required to obtain mutation resistance to DTG.
Table 2. Analysis of genetic barrier based on the minimum number of transitions and transversions required to obtain mutation resistance to DTG.
Codon
Position
SubstitutionSubtype C,
n (%) a
Subtype B,
n (%) b
Wild-Type CodonMutant CodonMinimal Score c
66T66A439 (96.91)1829 (97.08)ACAGTC, GCC/A/G1
3 (0.66)7 (0.37)ACG1
3 (0.66)6 (0.32)ACT2
8 (1.77)42 (2.23)ACC1
T66K439 (96.91)1829 (97.08)ACAAAA/G2.5
3 (0.66)7 (0.37)ACG2.5
3 (0.66)6 (0.32)ACT5
8 (1.77)42 (2.23)ACC5
T66I439 (96.91)1829 (97.08)ACAATT/C/A1
3 (0.66)7 (0.37)ACG3.5
3 (0.66)6 (0.32)ACT1
8 (1.77)42 (2.23)ACC1
92E92Q441 (97.35)446 (23.67)GAACAA/G2.5
12 (2,65%)1438 (76.33)GAG2.5
118G118R394 (86.98)1750 (92.89)GGCCGT/C/A/G, AGA/G2.5
19 (4.19)32 (1.7)GGA1
1 (0.22)5 (0.27)GGG1
39 (8.61)91 (4.83)GGT2.5
138E138A440 (93.16)1831 (97.19)GAAGTC, GCC/A/G2.5
13 (2.87)39 (2.07)GAG2.5
E138K440 (93.16)1831 (97.19)GAAGTC, GCC/A/G1
13 (2.87)39 (2.07)GAG1
E138T440 (93.16)1831 (97.19)GAAACT/C/A/G3.5
13 (2.87)39 (2.07)GAG3.5
140G140A243 (53.64)18 (0.96)GGGGTC, GCC/A/G2.5
208 (45.92)58 (3.08)GGA2.5
1 (0.22)201 (10.67)GGT3.5
1 (0.22)1607 (85.30)GGC2.5
G140S243 (53.64)18 (0.96)GGGTCT/C/A/G, AGT/C3.5
208 (45.92)58 (3.08)GGA3.5
1 (0.22)201 (10.67)GGT1
1 (0.22)1607 (85.30)GGC1
G140C243 (53.64)18 (0.96)GGGTGT, TTC5
208 (45.92)58 (3.08)GGA5
1 (0.22)201 (10.67)GGT2.5
1 (0.22)1607 (85.30)GGC3.5
143Y143C436 (96.25)1877 (99.63)TACTGT, TTC2.5
17 (3.75)7 (0.37)TAT1
Y143H436 (96.25)1877 (99.63)TACCAT/C1
17 (3.75)7 (0.37)TAT1
Y143R436 (96.25)1877 (99.63)TACCGT/C/A/G. AGA/G3.5
17 (3.75)7 (0.37)TAT2
147S147G403 (88.96)1828 (97.03)AGTGGT/C/A/G1
50 (11.04)56 (2.97)AGC1
148Q148H71 (15.67)1828 (97.03)CAACAT/C2.5
382 (84.33)56 (2.97)CAG2.5
Q148K71 (15.67)1828 (97.03)CAAAAA/G2.5
382 (84.33)56 (2.97)CAG2.5
Q148R71 (15.67)1828 (97.03)CAACGT/C/A/G, AGA/G1
382 (84.33)56 (2.97)CAG1
155N155H427 (94.26)1849 (98.14)AATCAT/C2.5
26 (5.74)35 (1.86)AAC2.5
263R263K62 (13.69)1833 (97.29)AGAAAA/G1
389 (85.87)4 (0.21)AGG1
a Subtype C: Ethiopian sequence used in the analysis (n = 453). b Subtype B: global subtype B sequence deposited before 2007 (before INSTI was used) retrieved from the Los Alamos database (n = 1884). c Minimal score calculated by the sum of number of transversions and transitions for each, with transitions scored as 1 and transversions scored as 2.5.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Arimide, D.A.; Szojka, Z.I.; Zealiyas, K.; Gebreegziabxier, A.; Adugna, F.; Sasinovich, S.; Björkman, P.; Medstrand, P. Pre-Treatment Integrase Inhibitor Resistance and Natural Polymorphisms among HIV-1 Subtype C Infected Patients in Ethiopia. Viruses 2022, 14, 729. https://doi.org/10.3390/v14040729

AMA Style

Arimide DA, Szojka ZI, Zealiyas K, Gebreegziabxier A, Adugna F, Sasinovich S, Björkman P, Medstrand P. Pre-Treatment Integrase Inhibitor Resistance and Natural Polymorphisms among HIV-1 Subtype C Infected Patients in Ethiopia. Viruses. 2022; 14(4):729. https://doi.org/10.3390/v14040729

Chicago/Turabian Style

Arimide, Dawit Assefa, Zsófia Ilona Szojka, Kidist Zealiyas, Atsbeha Gebreegziabxier, Fekadu Adugna, Sviataslau Sasinovich, Per Björkman, and Patrik Medstrand. 2022. "Pre-Treatment Integrase Inhibitor Resistance and Natural Polymorphisms among HIV-1 Subtype C Infected Patients in Ethiopia" Viruses 14, no. 4: 729. https://doi.org/10.3390/v14040729

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