Next Article in Journal
The Adenovirus Genome Contributes to the Structural Stability of the Virion
Next Article in Special Issue
Impact of Human Immunodeficiency Virus Type-1 Sequence Diversity on Antiretroviral Therapy Outcomes
Previous Article in Journal
Role of microRNAs in Arbovirus/Vector Interactions
Previous Article in Special Issue
Low-Frequency NNRTI-Resistant HIV-1 Variants and Relationship to Mutational Load in Antiretroviral-Naïve Subjects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Drug Resistance in Non-B Subtype HIV-1: Impact of HIV-1 Reverse Transcriptase Inhibitors

1
Christopher Bond Life Sciences Center, University of Missouri, Columbia, MO 65211, USA
2
Department of Molecular Microbiology and Immunology, University of Missouri, Columbia, MO 65211, USA
3
Division of Clinical Microbiology, Department of Laboratory Medicine, Karolinska Institute, Stockholm 141 86, Sweden
4
Clinical Research Center, Department of Infectious Diseases and Immunology, National Hospital Organization, Nagoya Medical Center, Nagoya 460-0001, Japan
5
Center for Advanced Biotechnology and Medicine, Rutgers University, Piscataway, NJ 08854, USA
6
Department of Chemistry and Chemical Biology, Rutgers University, Piscataway, NJ 08854, USA
7
Department of Oral and Craniofacial Science , School of Dentistry, University of Missouri, Kansas City, MO 64108, USA
8
Department of Microbiology and Molecular Genetics, University of Pittsburgh School of Medicine, Pittsburgh, PA 15219, USA
9
Department of Biochemistry, University of Missouri, Columbia, MO 65211, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Viruses 2014, 6(9), 3535-3562; https://doi.org/10.3390/v6093535
Submission received: 16 July 2014 / Revised: 9 September 2014 / Accepted: 9 September 2014 / Published: 24 September 2014
(This article belongs to the Special Issue HIV Drug Resistance)

Abstract

:
Human immunodeficiency virus (HIV) causes approximately 2.5 million new infections every year, and nearly 1.6 million patients succumb to HIV each year. Several factors, including cross-species transmission and error-prone replication have resulted in extraordinary genetic diversity of HIV groups. One of these groups, known as group M (main) contains nine subtypes (A-D, F-H and J-K) and causes ~95% of all HIV infections. Most reported data on susceptibility and resistance to anti-HIV therapies are from subtype B HIV infections, which are prevalent in developed countries but account for only ~12% of all global HIV infections, whereas non-B subtype HIV infections that account for ~88% of all HIV infections are prevalent primarily in low and middle-income countries. Although the treatments for subtype B infections are generally effective against non-B subtype infections, there are differences in response to therapies. Here, we review how polymorphisms, transmission efficiency of drug-resistant strains, and differences in genetic barrier for drug resistance can differentially alter the response to reverse transcriptase-targeting therapies in various subtypes.

1. Introduction

Approximately 35 million people are infected with different types and subtypes of HIV. Of the two major HIV types, HIV type 1 (HIV-1) causes ~95% of all infections, whereas HIV-2 (~5%) is mainly limited to West African countries (Senegal, Mali, Ghana, Guinea-Bissau, the Gambia, Niger, Cote d’Ivoire, and Liberia). HIV-1 has been divided into four groups: major (M), outlier (O), non‑M/non-O (N), and a new group P [1,2]. Group M accounts for more than 95% of all reported HIV‑1 infections [3].
Group M has been further divided into nine subtypes (also called clades): A-D, F-H, and J-K [2]. Of these, subtype C (HIV-1C) is the most prevalent worldwide and accounts for 50%–55% of all HIV infections [4]. It is mainly found in sub-Saharan Africa, Brazil, and India [4]. Subtype A (HIV-1A), which is approximately 20% of all HIV-1 infections, is primarily found in Eastern Europe, Northern Asia, and Western-Africa [4]. HIV-1 subtype B (HIV-1B) accounts for ~12% worldwide HIV-1 infections and is the main strain reported in developed countries (North America, Western Europe, Japan, and Australia). Other HIV-1 subtypes (D–K, or HIV-1D, HIV-1K, etc.) are rare (less than 6% of all infections). In addition to nine group M (HIV-M) subtypes, 66 circulating recombinant forms (CRFs) that evolve by recombination of two or more subtypes have been reported [5]. Although generally the geographic distribution of CRFs is not well defined, the CRF01_AE and CRF02_AG CRFs are predominately found in Southeast Asia and West Africa, respectively [6,7].
Most available drug-resistance data are derived from studies on HIV-1B, mainly because of the early availability of drugs and drug-resistance testing in developed countries. In contrast, similar data on non-B subtypes of HIV-1 (HIV-non-B) are limited, primarily due to historic delays in the availability of antiretroviral therapies (ART) in resource-limited countries. Increased access to ARTs, however, with significantly high non-adherence together with factors such as high rate of mutation, viral transmission and recombination in HIV-non-B infected patients have resulted in the emergence of drug-resistance data for other subtypes [8].
High genetic diversity of HIV-1 subtypes may influence the resistance pathways to currently used antiretrovirals (ARVs) [7]. The genetic differences can also impact the extent of cross-resistance to ARVs of the same class and lead to virologic failure, which may affect clinical outcome and immunological response [7,9]. Moreover, there is a considerable body of evidence that shows naturally occurring polymorphisms affect ARV response in subtype B virus [10,11,12,13,14,15,16,17]. A similar effect of polymorphisms can also be hypothesized for HIV-non-B viruses. In fact, recent reports have suggested that naturally occurring polymorphisms in HIV-non-B affect susceptibility to ARVs [18], although the impact of these polymorphisms on drug resistance is not well understood.
Current anti-HIV therapies, also known as highly active antiretroviral therapies (HAART), typically comprise combinations of nucleoside/nucleotide reverse transcriptase inhibitors (NRTIs), non-nucleoside reverse transcriptase inhibitors (NNRTIs), protease inhibitors (PIs), and integrase inhibitors (INSTIs). Generally, these ARVs are also effective against HIV-non-B viruses, even though they were originally designed and developed targeting HIV-1B virus. Nonetheless, susceptibility to ARVs can vary dramatically depending upon subtype due to inherent polymorphisms and cross-species transmission routes that may create differences in the drug resistance mechanisms of HIV-1 subtypes. Here, we summarize the status of HIV-non-B resistance to currently used ARTs in low- and middle-income countries, and present a comprehensive comparison with HIV-1B.

2. Resistance to Antiretroviral Therapies among Different HIV-1 Subtypes

Six different classes of drugs have been approved by US Food and Drug Administration (FDA) [19] for HIV-1 treatment. The two drug classes: (i) NRTIs; and (ii) NNRTIs target the viral enzyme reverse transcriptase (RT). RT is a multi-functional enzyme. It carries out both RNA- and DNA-dependent DNA polymerase activities, as well as RNase H activity (Figure 1). All three activities are required for the synthesis of dsDNA from viral ssRNA genome. Structures of HIV-1B RT in various functional and inhibited states have been reported over the past 22 years. These include structures in (i) unliganded form; (ii) in complex with various template-primers (T/P); (iii) T/P and dNTP; (iv) T/P and NRTI‑triphosphates (TP); (v) NNRTIs; or (vi) nucleic acid and NNRTIs [20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35].
Figure 1. The crystal structure of HIV-1 RT bound to double stranded DNA. HIV-1 RT is a heterodimer of p66 and p51 subunits. Due to the resemblance of p66 to a half-open right hand, different subdomains in p66 have been named as the “palm” (red), fingers (blue), and thumb (green) [33]. The polymerase and RNase H (orange) activities are located in p66 subunit, and are linked by the connection subdomain (yellow). The p51 (dark gray) subunit is derived from proteolytic cleavage of the RNase H from p66 and has identical primary and secondary structure. However, the tertiary structure of p51 is markedly different than p66, leading to a non-functional arrangement of catalytic residues. The template/primer (magenta/cyan) binds in a DNA-binding cleft formed primarily by the p66 subunit. The active site (marked as YMDD) and the NNRTI binding pocket (NNIBP) are shown as space-filled atoms. This figure was generated from Protein Data Bank file 2HMI [20]. This, as well as Figure 4 and Figure 5, were generated using PyMOL program [36].
Figure 1. The crystal structure of HIV-1 RT bound to double stranded DNA. HIV-1 RT is a heterodimer of p66 and p51 subunits. Due to the resemblance of p66 to a half-open right hand, different subdomains in p66 have been named as the “palm” (red), fingers (blue), and thumb (green) [33]. The polymerase and RNase H (orange) activities are located in p66 subunit, and are linked by the connection subdomain (yellow). The p51 (dark gray) subunit is derived from proteolytic cleavage of the RNase H from p66 and has identical primary and secondary structure. However, the tertiary structure of p51 is markedly different than p66, leading to a non-functional arrangement of catalytic residues. The template/primer (magenta/cyan) binds in a DNA-binding cleft formed primarily by the p66 subunit. The active site (marked as YMDD) and the NNRTI binding pocket (NNIBP) are shown as space-filled atoms. This figure was generated from Protein Data Bank file 2HMI [20]. This, as well as Figure 4 and Figure 5, were generated using PyMOL program [36].
Viruses 06 03535 g001
Structure(s) of HIV-non-B RTs have not been determined, although a structure of an unliganded HIV-2 RT has been reported [37]. However, due to the high sequence conservation between HIV-1B and HIV-non-B RTs, the overall folding of HIV-non-B RT is expected to be similar. RT is an asymmetric heterodimer comprising two structurally distinct subunits (p66 and p51). The p66 subunit contains the active sites for the polymerase and RNase H activities of the enzyme. The p51 subunit, which is derived from p66 by protease-mediated cleavage of the RNase H domain plays a structural role [26,32,33], although some mutations in p51 have been reported to affect catalytic activity as well [38,39].
The first anti-HIV drugs were nucleoside analogs lacking the 3'-OH group, which is required for DNA synthesis [40]. Currently, all approved NRTIs lack a 3'-OH (Figure 2), are phosphorylated in cells into dNTP analogs that bind at the dNTP-binding site [21,23,41,42], and act as chain terminators after their incorporation into viral DNA by RT [40,43,44].
The second class of anti-HIV-1 RT drugs, (NNRTIs, Figure 3), bind in a hydrophobic pocket of HIV-1B RT at the base of the p66 thumb subdomain, ~10 Å away from the polymerase active site (Figure 1). NNRTI binding has multiple effects, including restriction of the p66 thumb subdomain’s flexibility, and repositioning of the nucleic acid [25,26,33,43,45,46].

2.1. Mechanism of NRTI Resistance in HIV-1B

In HIV-1B, major mutations associated with NRTI resistance are M41L, A62V, D67N, 69 insertions (T69S plus the addition of SS, SA, or SG), K70R/E, K65R, L74V, V75I, F77L, Y115F, F116Y, Q151M, M184V/I, L210W, T215Y/F, and K219Q/E (Figure 4) [47]. Of these, M41L, A62V, 69ins, K70R, L210W, T215Y/F, and K219Q/E form the 69 insertion complex. Mutations A62V, V75I, F77L, F116Y, and Q151M combine to form the Q151 complex, and mutations M41L, D67N, K70R, L210W, T215Y/F and K219Q/E are known as TAMs (thymidine-associated mutations) or EEMs (excision-enhancing mutations) [47]. The 69 insertion complex is associated with resistance to all approved NRTIs when present with one or more TAMs at codons 41, 210, or 215 [47,48]. The Q151 complex mutations confer reduced susceptibility to all NRTIs except TDF (tenofovir) [48] while the TAMs enhance resistance to all approved NRTIs [49,50].
Figure 2. FDA approved nucleoside/nucleotide inhibitors (NRTIs/NtRTIs) of HIV-1B RT: All approved NRTIs/NtRTIs (AZT, azidothymidine; d4T, stavudine; FTC, emtricitabine; ddC, zalcitabine; ddI, didanosine; ABC, abacavir; TDF, tenofovir) lack a 3'-OH group. After incorporation of NRTIs/NtRTIs into DNA by HIV-1B RT, they act as chain terminators. FTC is 5-fluoro derivative of 3TC; the latter is not included in this figure. All chemical structures in this article were drawn by ChemSketch 3.5 [51].
Figure 2. FDA approved nucleoside/nucleotide inhibitors (NRTIs/NtRTIs) of HIV-1B RT: All approved NRTIs/NtRTIs (AZT, azidothymidine; d4T, stavudine; FTC, emtricitabine; ddC, zalcitabine; ddI, didanosine; ABC, abacavir; TDF, tenofovir) lack a 3'-OH group. After incorporation of NRTIs/NtRTIs into DNA by HIV-1B RT, they act as chain terminators. FTC is 5-fluoro derivative of 3TC; the latter is not included in this figure. All chemical structures in this article were drawn by ChemSketch 3.5 [51].
Viruses 06 03535 g002
Figure 3. FDA-approved NNRTIs: nevirapine (NVP), efavirenz (EFV), delavirdine (DLV), etravirine (ETV), and rilpivirine (RPV). Currently, DLV is not recommended by the US Department of Health and Human Services. ETV, RPV, and sometimes EFV, are referred to as second-generation NNRTIs.
Figure 3. FDA-approved NNRTIs: nevirapine (NVP), efavirenz (EFV), delavirdine (DLV), etravirine (ETV), and rilpivirine (RPV). Currently, DLV is not recommended by the US Department of Health and Human Services. ETV, RPV, and sometimes EFV, are referred to as second-generation NNRTIs.
Viruses 06 03535 g003
Figure 4. Stereoview of the polymerase active site of HIV-1B RT with residues involved in NRTI resistance. Amino acid residues, dNTP and template substrates are shown in ball-and-stick representation. The backbone of HIV-1 RT (PDB file 1RTD) [21] is shown in light gray ribbon representation. The amino acid side chains are colored by atom type (carbon, white; oxygen, red; nitrogen, blue; and sulfur, yellow). The dNTP is also colored by atom type but the carbon and phosphorus atoms are green and orange, respectively. The template nucleotide is colored magenta. For clarity, the position of F77 is not shown.
Figure 4. Stereoview of the polymerase active site of HIV-1B RT with residues involved in NRTI resistance. Amino acid residues, dNTP and template substrates are shown in ball-and-stick representation. The backbone of HIV-1 RT (PDB file 1RTD) [21] is shown in light gray ribbon representation. The amino acid side chains are colored by atom type (carbon, white; oxygen, red; nitrogen, blue; and sulfur, yellow). The dNTP is also colored by atom type but the carbon and phosphorus atoms are green and orange, respectively. The template nucleotide is colored magenta. For clarity, the position of F77 is not shown.
Viruses 06 03535 g004
Mutations K65R, L74V, Y115F, and M184V are associated with resistance to abacavir (ABC), K65R and L74V are ddI-resistance mutations, K65R and M184V/I are resistant to FTC and 3TC, K65R and K70E are related to TDF resistance, and M41L, D67N, K70R, L210W, T215Y/F, and K219Q/E cause resistance to AZT [47].
Resistance to NRTIs is achieved by two distinct mechanisms in HIV-1B: (i) discrimination; and (ii) excision. In the discrimination mechanism, the mutations favor natural dNTP substrate over NRTI/NtRTI by decreasing binding affinity of NRTI-TP (Kd), by reducing polymerization rate (kpol), or both [44,52,53,54,55,56]. The mutations known to cause resistance using the discrimination mechanism are K65R, K70E, L74V, V75I, Q151M complex, and M184V/I [52,53,55,57,58,59,60,61,62,63,64,65,66,67]. Biochemical data show that most of these mutations reduce the polymerase rate (kpol) and/or decrease the NRTI-TP binding affinity (Kd) compared to the wild-type HIV-1B RT [40,43,52,53,55,57,58,59,60,61,62,63,64,65,66,67].
Using the excision mechanism, an EEM HIV-1B RT carries out ATP-mediated excision of NRTI from the 3'-end of the DNA primer chain and permits extension of the DNA [57,68,69]. This process is also referred to as “primer unblocking”. AZT and d4T resistance mutations M41L, D67N, K70R, L210W, T215Y/F, and K219Q/E are the most common examples of EEMs. These residues are located outside the dNTP-binding site [21,23,24,40,43,70,71,72] and do not affect the incorporation of nucleotide analogs [56,73,74,75,76,77]. Instead, they facilitate the ATP-dependent removal of the incorporated NRTI-MP that is catalyzed by the same active site as the polymerization reaction but in the reverse direction [69,70,78,79,80,81,82,83].

2.2. NRTI-Associated Mutations in Different HIV-1 Subtypes

A disparity in resistance mutations after NRTI treatment has been seen among HIV-1 subtypes. HIV-1C patients from Botswana who underwent AZT/ddI-containing HAART developed resistance predominantly through the 67N/70R/215Y pathway [84]. These mutations are different than AZT/ddI-associated resistance mutations (TAM/EEM) in HIV-1B. In HIV-1B, AZT/ddI resistance is acquired through either a TAM1 (M41L, L210W and T215Y) or a TAM2 (D67N, K70R and K219E/Q) pathway. A different set of mutations was selected in HIV-1C infected patients in India [85] under the same treatment protocol. While most of the NRTI-resistance mutations in the Indian patients were similar to those in HIV-1B (TAM1 and TAM2), they also harbored additional RT mutations A98G, E203D/K/V/N/A, H208Y, and H221Y [85]. Similarly, an entirely different resistance pattern was seen in HIV-1C infected patients from Malawi and South Africa who received first-line treatment with d4T/3TC/NVP [86,87]. Not surprisingly, the most prevalent NRTI-associated mutation in these patients was M184V (81%), which is associated with 3TC treatment. Interestingly, 23% of the patients acquired the K70E or K65R mutations that are associated with TDF resistance even though the patients were not exposed to TDF [86]. Results from another study in Botswana also showed high incidences of the K65R mutation (30%) in HIV-1C patients who received combinations of either d4T/ddI/NVP or d4T/ddI/EFV [88]. The K65R mutation was detected in 7% and 15% of patients in South Africa failing first- and second-line regimens, respectively [89,90,91]. The nucleoside backbone in these treatments included d4T/3TC or ddI/AZT [90,91]. The K65R mutation in HIV-1C was also seen in Ethiopian and Indian patients (10%–12%) who failed d4T/3TC/NVP therapy [92,93]. These results suggest that the K65R substitution may emerge at a higher frequency among individuals infected with HIV-1C viruses who are receive d4T/3TC-based first-line therapy, and the introduction of TDF-based regimens may show limited success [9].
The K65R mutation has been shown to be selected faster in HIV-1C than in HIV-1B when tenofovir (TDF) is used as one of the nucleoside drugs [7]. On the contrary, K65R appears to be less frequent in HIV-1A than other subtypes [94]. Some subtype-dependent differences in K65R or TAM mutations can be attributed to treatment regimens, disease stage and access to viral load testing in different developing countries [94]. The faster acquisition of K65R in HIV-1C has also been linked to the nature of the HIV-1CRNA template [95,96,97].

2.3. NNRTI-Associated Mutations in Different HIV-1 Subtypes

EFV and NVP are two NNRTIs that have been widely used in both developed and resource-limited countries. However, both NNRTIs have a low genetic barrier [98]. A single mutation in the NNRTI binding pocket is sufficient to cause clinical failure [98]. Moreover, a high level of cross-resistance between NVP and EFV [99] has limited available therapeutic options. ETV and RPV are second generation NNRTIs that are effective against common NVP and/or EFV resistant HIV-1 viruses [100,101,102]. NNRTIs bind to a hydrophobic pocket (non-nucleoside inhibitor binding pocket or NNIBP) formed by residues 95, 100, 101, 103, 106, 108, 179, 181, 188, 190, 227, 229, 234, 236, and 318 from the p66 subunit and 138 from the p51 subunit (Figure 5) [27,30,45]. NNRTI resistance can be achieved by two different mechanisms: (i) affecting the entry/release (kon/koff) of NNRTI to and from the NNIBP (e.g., K103N and E138K mutations) [38,103,104,105,106,107]; and (ii) altering the NNIBP geometry, thus directly affecting interactions with the inhibitor (e.g., Y181C and Y188L) [27,29,72,108,109].
The NNRTIs are used by themselves and in combination with two NRTIs. Similar to diverse mutation pattern in response to NRTI treatment, different patterns of NNRTI-resistance mutations have been reported for different HIV subtypes. Single-dose NVP treatment, often used for prevention of mother-to-child transmission (PMTCT) in countries with limited resources, has shown a wide range of subtype-dependent resistance patterns. Resistance to NVP monotherapy in HIV-1C, HIV-1D, HIV1-A and HIV-1 CRF02_AG infections has been reported to be 69%, 36%, 19%, and 21%, respectively [110]. Similarly, a meta-analysis in South African HIV-1C patients treated for PMTCT showed that 35.7% of women and 52.6% of children carried virus with NVP-resistance mutations [111]. Results from a Ugandan cohort showed that HIV-1C was more NVP resistant than HIV-1D and HIV-1A [112]. The K103N mutation most frequently appeared in mothers, while the Y181C mutation was predominant in children [113]. These results clearly demonstrate that the response to NVP monotherapy (as for PMTCT) significantly varies among HIV subtypes.
During in vitro serial passages of HIV-1C in NVP or EFV, the virus developed the V106M mutation, in addition to the V106A mutation, which is commonly selected in HIV-1B [114]. This difference has been attributed to a nucleotide polymorphism at position 106 of RT [18]. The clinical importance of V106M in HIV-non-B subtypes has been confirmed in many studies, which show that V106M emerges frequently in patients infected with HIV-1C and HIV-1 CRF01_AE viruses during therapy with NVP or EFV [80,84,85,115,116,117,118]. It should be noted that, similar to residue 106, additional silent drug-resistant RT mutations at residues 65, 138, and 161 were reported in Ethiopian isolates and subtype C reference strains [114].
Figure 5. Amino acids involved in HIV-1B NNRTI resistance: Stereoview of HIV-1B RT with RPV bound at the NNIBP and NNRTI-resistance residues and RPV shown in ball-and-stick representation. The RT backbone (PDB file 2ZD1) [30] is shown in ribbon representation (p66 in light gray and p51 in magenta). Amino acid side chains are colored with carbon in white, oxygen in red, nitrogen in blue, and sulfur in yellow. RPV carbons are shown in green and nitrogens in blue. RPV resistance residue E138 [103,104,119,120,121] is shown in magenta. Residues 227, 229, and 234 are omitted for clarity.
Figure 5. Amino acids involved in HIV-1B NNRTI resistance: Stereoview of HIV-1B RT with RPV bound at the NNIBP and NNRTI-resistance residues and RPV shown in ball-and-stick representation. The RT backbone (PDB file 2ZD1) [30] is shown in ribbon representation (p66 in light gray and p51 in magenta). Amino acid side chains are colored with carbon in white, oxygen in red, nitrogen in blue, and sulfur in yellow. RPV carbons are shown in green and nitrogens in blue. RPV resistance residue E138 [103,104,119,120,121] is shown in magenta. Residues 227, 229, and 234 are omitted for clarity.
Viruses 06 03535 g005
Data from the DUET-1 and DUET-2 phase III clinical trials revealed a total of 17 ETV resistance-associated mutations (RAMs) (V90I, A98G, L100I, K101E/H/P, V106I, E138A, V179D/F/T, Y181C/I/V, G190A/S, and M230L) [100,101]. Three or more of these mutations were associated with decreased virological response to ETV [122]. V90I and E138A were the most prevalent mutations in HIV-1 CRF02_AG and HIV-1A1, while V106I was most prevalent in HIV-1C and HIV-1D. Biochemical data showed that E138A in combination with V179E and Y181C had increased resistance to ETV by five-fold and 11-fold, respectively [122]; however, none of the dual mutation samples were found to be ETV resistant [122].
In vitro serial passages of HIV-1B, HIV-1C, and HIV-1 CRF02_AG in ETV showed distinct patterns of resistance mutations [123]. HIV-1C acquired the V90I, V106M, E138K, V179D/E, Y181C, G190E/G, M230L, and K238N mutations. These mutations also emerged in HIV-1 CRF02_AG. However, in addition to the above mutations (except for M230L), ETV also selected K101Q, V189I, and H221Y in HIV-1B. Interestingly, the K103N/Q mutations were selected in HIV-1B and HIV-1 CRF02_AG, but not in HIV-1C.
In low- and middle- income countries (LMICs) where phenotyping and genotyping is not widely available and where non-B strains dominate, the patients’ failure to a treatment regimen is always monitored either by immunological criteria (CD4+ T-cell counts) or manifestation of clinical symptoms. In such cases, extensive cross-resistance is observed, even when NVP and EFV are replaced with the second generation NNRTI etravirine (ETV). A study from India reported that in a failing regimen including NVP or EFV, 40% of the HIV-1C infected patients had at least partial resistance to ETV [124]. Similar observations were reported in another Indian study where the HIV-1C-infected patients were treated with NVP/EFV for one to eight years and 45% were resistant to ETV [93].
A group of 17 mutations, L100I, K101E/P, E138A/G/K/Q/R, V179L, Y181C/I/V, Y188L, H221Y, F227C, and M230I/L, have been defined as RPV RAMs [47,125,126,127,128]. The prevalence of these mutations among RPV treatment-naïve (but treated with NRTI+PI or NRTI+NNRTI regimens) patients infected with HIV-1B and HIV-non-B has been reported [125]. Nearly 5% of the patients harbored primary RPV mutations. E138A was the most prevalent mutation (3.2%) followed by E138K, H221Y, E138G, Y181C, and Y188L in HIV-1C. The distribution of RPV RAMs was significantly different between HIV-1B (2%) and HIV-non-B (5%) subtype patients [125]. This difference can be attributed to E138A, which is a known polymorphism mutation (6%–8%) in HIV-1C virus [129].
Despite the structural similarity between ETV and RPV, both being DAPY (diarylpyrimidine) derivatives, different sets of resistance mutations have been reported for the same HIV-non-B (HIV-CRF01_AE) infection. The mutations Y181C, G190A, and K103N emerged in patients who failed ETV therapy, whereas mutations K101P, Y181I, and Y181V were predominantly seen in patients who failed RPV [130]. Notably, the patients who failed RPV therapy were resistant to ETV even though ETV RAMs were not found among these patients [125,130]. Many other NNRTI-resistance mutations that vary from subtype to subtype have been reported. A list of these mutations has been is given in Table 1. In some subtypes, resistance to given NNRTI(s) has been attributed to a high degree of polymorphism. For example, A98S mutation, which is resistant to all NNRTIs in HIV-1B and HIV-1C (Table 1) may confer resistance to NNRTIs in HIV-1G due to high polymorphism in this subtype [131].
Table 1. NNRTI-resistance mutations in response to NNRTI treatment among different HIV subtypes (modified from Santoro and Perno, [132]).
Table 1. NNRTI-resistance mutations in response to NNRTI treatment among different HIV subtypes (modified from Santoro and Perno, [132]).
MutationSubtypeNNRTIsCommentReference
K103NB, C, F, CRF02_AEEFV, DLV, NVPK103 appears with low frequency in C compared to B, F and CRF02_AE[133,134,135,136]
V106MB, C and CRF01_AEEFV, NVPV106M emerges frequently in C due to low genetic barrier compared to B and CRF01_AE[99,114,118]
E138KB, C, CRF02_AGETRFirst emerges in B, C and CRF02_AG. Preferential selection in B is Y181C[80,123]
G190ACAll NNRTIsThis mutation in C may be due to high G190A/S polymorphism in C[118]
Y181CA, BETRPreferential selection under drug pressure on A and B subtypes[80]
Y181C and Y188LCEFV, DLV, NPVHigh frequency in C[133,134,135,136]
N348ICETRHigh frequency in C at ETR failure[137,138]
N348IBNVPHigh frequency in B at NVP failure[11,12]

2.4. Mutations that Impact both Classes of RTIs (Connection Subdomain Mutations)

Until recently, all HIV-1 drug-resistance testing kits interrogated ~250 N-terminal amino acids of HIV-1 RT. These testing kits were based on the assumption that all currently approved RTIs target the polymerase domain of RT, and mutations in the domain confer drug resistance. One group of researchers from Japan, in collaboration with us [11] and another group from Australia [12], independently showed that the connection subdomain mutation N348I causes resistance to both classes of drugs (NRTIs and NNRTIs) in HIV-1B and HIV-1D patients. Biochemical data showed that N348I, by itself or in combination with TAMs, reduces the rate of RNA template degradation by RT [12,139]. We used transient-state kinetics to decipher the contribution of N348I in the polymerase function of RT and resistance to NVP [38]. We demonstrated that the N348I mutation increases the processivity of RT and enhances NVP dissociation from the NNIBP [38]. A molecular dynamics modeling study proposes that N348I contributes to NVP resistance through a long-range allosteric communication network between the connection and NNIBP [140]. More recently, Brehm et al. [138] reported the presence of the N348I mutation in HIV-1C patients who failed d4T/3TC/NVP or d4T/3TC/EFV therapy. A group of drug-resistance mutations in the connection subdomain of RT (A376S, A400T, Q334D, G335D, N348I, and A371V) in a Brazilian cohort (HIV-1C) was recently identified using single genome sequencing methodology [141]. The connection subdomain mutations identified in this study were similar to those previously reported in subtype B. However, the data showed that these mutations enhanced resistance to ETV by six- to 11-fold in the presence of L100I/K103N/Y181C (NNRTI) and TAM-associated mutations [141]. A detailed kinetic characterization of connection subdomain mutations in HIV-non-B needs to be performed before one can establish whether their mechanism of inhibition and resistance is similar to that in HIV-1B.

3. Transmitted Drug Resistance (TDR) in HIV-Non-B

Transmitted drug resistance (TDR) is defined as resistance to one or more ARVs in individuals with no previous drug exposure, and it is presumed to be the result of the direct transmission of resistant strains from treated individuals [43]. TDR can be complicated by the presence of drug resistance mutations in a given subtype that may exist as polymorphisms in some other subtype. Both TDR and polymorphisms can lead to ART failure in treatment-naïve patients and limit the efficacy of drug regimens. In regions of the world where HAART has been used for several years, TDR can be found in up to 25% of HIV infected individuals [128,142,143,144]. The most commonly reported NRTI TDR mutations in HIV-1B are M184V and M184I. Other TDR mutations are M41L, K219E/N/Q/R, and T215F/Y and 215 revertant mutations (T215D/C/E/S/I/V) [145,146]. Transmission of M184V has also been reported in HIV-non-B [143] cases with high-level resistance to 3TC and FTC. The most common HIV-non-B TDR mutations are K103N, Y181C/I, and Y188C/H/L [143,144,147,148].

4. Polymorphisms in HIV-Non-B

A polymorphism is a nucleotide variation from the consensus sequence within a subtype, arbitrarily defined as a position-specific difference occurring in >1% of sequences [147]. Biochemical and virological data have demonstrated that polymorphisms can affect the magnitude of resistance conferred by known resistance mutations [10,11,15,148,149], not only in different HIV-1 subtypes but also in different HIV types. For example, HIV-2 and group O viruses have high-level innate resistance to NNRTIs because of drug-resistance polymorphisms [150,151]. Of the first 240 amino acid residues of HIV-1 RT, polymorphisms at 41 positions in HIV-1A (17%), 45 in HIV-1B (19%), 56 in HIV-1C (23%), 43 in HIV-1D (18%), 35 in HIV-1F (15%), 32 in HIV-1G (13%), 46 in HIV-1 CRF01_AE (19%), and 56 in HIV-1 CRF021_AG (23%) have been reported based on testing of drug-naïve HIV-infected individuals [131,147]. Subtype-specific polymorphisms at residue positions that are related to NRTI and NNRTI drug resistance are shown in Table 2.
Table 2. Polymorphism at drug resistance positions in RT from untreated patients (modified from Kantor and Katzenstein, [147]).
Table 2. Polymorphism at drug resistance positions in RT from untreated patients (modified from Kantor and Katzenstein, [147]).
SubtypeSubtype Reference SequencesRT Sequence PM Position, NRTI RAMsRT Sequence PM Position, NNRTI RAMs
AU455A62V, V118IE138A, V179I/T, K238R
BHXB2V118IA98S, V179I/D/E
CC2220V118IA98S, E138A, V179I
DNDKV118I (comparable to B)V179I
F93BR020V41I, M184VE138A/G, V179I/D
GSE6165 A98S, V179E
CRF01_AECM240V41I, A62VV75L, V106I, V179I/D
CRF02_AGIbNGV41I, V118IV106I, V179I
It is now well established that connection subdomain mutations are associated with low-level resistance to NNRTIs [11,12,152]. Several connection subdomain mutations are polymorphic in HIV-non-B [153,154,155]. Polymorphisms G335D and A371V in HIV-1 CRF01_AE showed significantly reduced susceptibility to NRTIs and NNRTIs [154]. In a Brazilian HIV-1C cohort, polymorphisms G335D, A376S, and A400T/S have been reported [153]. However, more data from different HIV-non-B subtypes are required to establish the effect of polymorphisms on drug susceptibility and drug resistance.

5. Therapeutic Response in HIV-Non-B subtypes in the Presence of Drug Resistance Mutations

Several clinical studies have provided evidence of viral re-suppression despite the presence of major drug-resistance mutations (DRMs) in HIV-1C. A South African study showed that despite the presence of key NNRTI mutations like K103N, V106M, or G190A, a portion of patients achieved viral re-suppression [156]. This was further supported by a report from another South African cohort where, in nine of 12 viremic patients, HIV-1C was suppressed even in the presence of major DRMs, such as K103N or M184V [157]. A recent study from India, where HIV-1C is the prevalent subtype, reported that three patients had controlled viremia in the presence of multiple major NRTI and NNRTI resistance-associated mutations (M184V+G190A; M184V+K103N; M184V+Y181C) without change in therapy; the patients maintained >95% adherence both before and after the viral rebound as well as CD4+ T-cell counts in the normal range [158]. Collectively, these studies support the idea that further work is needed to understand how DRMs phenotypically affect viral fitness, drug susceptibility, and patients’ therapeutic responses.

6. Future Perspectives: New Class of NRTIs

The combination of NRTIs currently in use can limit treatment efficacy in many ways [159]. For example, AZT and d4T share the same excision pathway and therefore combining the two drugs can produce suboptimal results. Similarly, drug-drug interactions such as those reported between abacavir (ABC) and other protease inhibitors (atazanavir/ritonavir or lopinavir/ritonavir) may also impact the outcome of some combination therapies [160]. The low genetic barrier of NNRTIs, higher polymorphism in NNIBP of HIV-non-B subtypes, and extensive cross-resistance to second generation NNRTIs after failure to first generation NNRTIs may present challenges for use of NNRTI-containing combinations in some cases. NRTIs that act by novel inhibition mechanisms, which can be used synergistically with current drugs to improve resistance profiles, are continuously pursued. The chemical structure of some new NRTIs, which are either in pre-clinical development or at various stages of clinical trials, are shown in Figure 6. Amdoxovir is a prodrug of β-d-dioxolane guanosine (DXG), currently evaluated in phase II clinical trials. Apricitabine (ATC) is in phase IIb clinical trials and it has been shown to be effective against HIV-1 viruses containing M184V, L74V, and TAMs [161].
Several 4'-substituted NRTIs have been reported to be potent antivirals [103,162,163]. Among them, 4'-ethynyl-2-fluoro-2'-deoxyadenosine (EFdA) (Figure 6) is the most potent that can inhibit RT as a translocation-defective RT inhibitor or as a delayed chain terminator, depending on the sequence of the template/primer substrate [164,165]. These compounds contain an ethynyl group modification at the 4' position of the sugar moiety that appears to make extensive interactions in a hydrophobic pocket of the RT polymerase active site and block the translocation of RT after their incorporation into the nascent DNA chain [163]. EFdA inhibits efficiently essentially all drug-resistant strains [166], and displays low mitochondrial toxicity due to the fact that it is a poor substrate for DNA polymerase γ [164,167]. The effect of these new NRTIs on HIV-non-B subtypes is yet to be determined.
Figure 6. NRTIs in clinical trials or pre-clinical development. Amdoxovir, apricitabine, and elvucitabine are being developed by RFS Pharma, Avexa Pharmaceuticals, and Achillion Pharmaceuticals, respectively. EFdA has been licensed by Merck from Yamasa Corporation. With the exception of EFdA, these compounds lack a 3'-OH and function as chain terminators. EFdA contains a 3'-OH and blocks viral replication by multiple mechanisms, including by acting as a translocation-deficient reverse transcriptase inhibitor [103,163,164,168].
Figure 6. NRTIs in clinical trials or pre-clinical development. Amdoxovir, apricitabine, and elvucitabine are being developed by RFS Pharma, Avexa Pharmaceuticals, and Achillion Pharmaceuticals, respectively. EFdA has been licensed by Merck from Yamasa Corporation. With the exception of EFdA, these compounds lack a 3'-OH and function as chain terminators. EFdA contains a 3'-OH and blocks viral replication by multiple mechanisms, including by acting as a translocation-deficient reverse transcriptase inhibitor [103,163,164,168].
Viruses 06 03535 g006

7. Conclusions

In summary, the available biochemical, virological and clinical data on HIV-1B and HIV-non-B indicate that the genetic diversity among HIV-1 subtypes can impact the outcome of ART. As HAART becomes more accessible to more patients in resource-limited settings, a better picture of susceptibility and resistance is expected to emerge. These results may influence (i) the formulation of currently approved drugs and drug combinations to improve efficacy and adherence in HIV-non-B infected patients and (ii) the development of more potent and broad-spectrum NRTIs and NNRTIs.

Acknowledgments

This work was supported, in whole or in part, by National Institutes of Health grants AI076119, AI099284, AI100890, AI112417, and GM103368 (SGS); R37 AI27690 (MERIT Award) and P50 GM103368 to EA; AI079801 to (MAP) and Swedish International Developing Agency, CHAIN FP7 EU, the Swedish Civil Contingencies Agency SWE-2009-151, and the Swedish Research Council 521-2012-3476 and 2007-7092 to AS. We also acknowledge support from Ministry of Knowledge and Economy, Bilateral International Collaborative R&D Program, Republic of Korea and the financial support from Mizzou Advantage program of University of Missouri (SGS).

Conflicts of Interest

Authors declare no conflict of interest.

References and Notes

  1. Plantier, J.C.; Leoz, M.; Dickerson, J.E.; de Oliveira, F.; Cordonnier, F.; Lemee, V.; Damond, F.; Robertson, D.L.; Simon, F. A new human immunodeficiency virus derived from gorillas. Nat. Med. 2009, 15, 871–872. [Google Scholar]
  2. Robertson, D.L.; Anderson, J.P.; Bradac, J.A.; Carr, J.K.; Foley, B.; Funkhouser, R.K.; Gao, F.; Hahn, B.H.; Kalish, M.L.; Kuiken, C.; et al. HIV-1 nomenclature proposal. Science 2000, 288, 55–56. [Google Scholar]
  3. Ward, M.J.; Lycett, S.J.; Kalish, M.L.; Rambaut, A.; Leigh Brown, A.J. Estimating the rate of intersubtype recombination in early HIV-1 group M strains. J. Virol. 2013, 87, 1967–1973. [Google Scholar] [CrossRef]
  4. Hemelaar, J.; Gouws, E.; Ghys, P.D.; Osmanov, S.; WHO-UNAIDS Network for HIV Isolation and Characterisation. Global trends in molecular epidemiology of HIV-1 during 2000–2007. AIDS 2011, 25, 679–689. [Google Scholar]
  5. Los Alamos National Laboratory. HIV Circulating Recombinant Forms (CRFs). Available online: http://www.hiv.lanl.gov/content/sequence/HIV/CRFs/CRFs.html (accessed on 15 September 2014).
  6. Vidal, N.; Peeters, M.; Mulanga-Kabeya, C.; Nzilambi, N.; Robertson, D.; Ilunga, W.; Sema, H.; Tshimanga, K.; Bongo, B.; Delaporte, E. Unprecedented degree of human immunodeficiency virus type 1 (HIV-1) group M genetic diversity in the Democratic Republic of Congo suggests that the HIV-1 pandemic originated in Central Africa. J. Virol. 2000, 74, 10498–10507. [Google Scholar]
  7. Brenner, B.G. Resistance and viral subtypes: How important are the differences and why do they occur? Curr. Opin. HIV AIDS 2007, 2, 94–102. [Google Scholar] [CrossRef]
  8. Phillips, A.N.; Dunn, D.; Sabin, C.; Pozniak, A.; Matthias, R.; Geretti, A.M.; Clarke, J.; Churchill, D.; Williams, I.; Hill, T.; et al. Long term probability of detection of HIV-1 drug resistance after starting antiretroviral therapy in routine clinical practice. AIDS 2005, 19, 487–494. [Google Scholar]
  9. Wainberg, M.A.; Brenner, B.G. Role of HIV Subtype Diversity in the Development of Resistance to Antiviral Drugs. Viruses 2010, 2, 2493–2508. [Google Scholar] [CrossRef]
  10. Hachiya, A.; Marchand, B.; Kirby, K.A.; Michailidis, E.; Tu, X.; Palczewski, K.; Ong, Y.T.; Li, Z.; Griffin, D.T.; Schuckmann, M.M.; et al. HIV-1 Reverse Transcriptase (RT) Polymorphism 172K Suppresses the Effect of Clinically Relevant Drug Resistance Mutations to Both Nucleoside and Non-nucleoside RT Inhibitors. J. Biol. Chem. 2012, 287, 29988–29999. [Google Scholar]
  11. Hachiya, A.; Kodama, E.N.; Sarafianos, S.G.; Schuckmann, M.M.; Sakagami, Y.; Matsuoka, M.; Takiguchi, M.; Gatanaga, H.; Oka, S. Amino acid mutation N348I in the connection subdomain of human immunodeficiency virus type 1 reverse transcriptase confers multiclass resistance to nucleoside and nonnucleoside reverse transcriptase inhibitors. J. Virol. 2008, 82, 3261–3270. [Google Scholar] [CrossRef]
  12. Yap, S.H.; Sheen, C.W.; Fahey, J.; Zanin, M.; Tyssen, D.; Lima, V.D.; Wynhoven, B.; Kuiper, M.; Sluis-Cremer, N.; Harrigan, P.R.; et al. N348I in the connection domain of HIV-1 reverse transcriptase confers zidovudine and nevirapine resistance. PLoS Med. 2007, 4, e335. [Google Scholar] [CrossRef] [Green Version]
  13. Betancor, G.; Puertas, M.C.; Nevot, M.; Garriga, C.; Martinez, M.A.; Martinez-Picado, J.; Menendez-Arias, L. Mechanisms involved in the selection of HIV-1 reverse transcriptase thumb subdomain polymorphisms associated with nucleoside analogue therapy failure. Antimicrob. Agents Chemother. 2010, 54, 4799–4811. [Google Scholar] [CrossRef]
  14. Wright, D.W.; Deuzing, I.P.; Flandre, P.; van den Eede, P.; Govaert, M.; Setiawan, L.; Coveney, P.V.; Marcelin, A.G.; Calvez, V.; Boucher, C.A.; et al. A polymorphism at position 400 in the connection subdomain of HIV-1 reverse transcriptase affects sensitivity to NNRTIs and RNaseH activity. PLoS One 2013, 8, e74078. [Google Scholar]
  15. Kearney, M.; Palmer, S.; Maldarelli, F.; Shao, W.; Polis, M.A.; Mican, J.; Rock-Kress, D.; Margolick, J.B.; Coffin, J.M.; Mellors, J.W. Frequent polymorphism at drug resistance sites in HIV-1 protease and reverse transcriptase. AIDS 2008, 22, 497–501. [Google Scholar] [CrossRef]
  16. Kemp, S.D.; Shi, C.; Bloor, S.; Harrigan, P.R.; Mellors, J.W.; Larder, B.A. A novel polymorphism at codon 333 of human immunodeficiency virus type 1 reverse transcriptase can facilitate dual resistance to zidovudine and L-2',3'-dideoxy-3'-thiacytidine. J. Virol. 1998, 72, 5093–5098. [Google Scholar]
  17. Delviks-Frankenberry, K.A.; Nikolenko, G.N.; Pathak, V.K. The “Connection” between HIV drug resistance and RNase H. Viruses 2010, 2, 1476–1503. [Google Scholar] [CrossRef]
  18. Wainberg, M.A.; Brenner, B.G. The impact of HIV genetic polymorphisms and subtype differences on the occurrence of resistance to antiretroviral drugs. Mol. Biol. Int. 2012, 2012, 256982. [Google Scholar] [CrossRef]
  19. Chen, R.; Quinones-Mateu, M.E.; Mansky, L.M. Drug resistance, virus fitness and HIV-1 mutagenesis. Curr. Pharm. Des. 2004, 10, 4065–4070. [Google Scholar] [CrossRef]
  20. Ding, J.; Das, K.; Hsiou, Y.; Sarafianos, S.G.; Clark, A.D., Jr.; Jacobo-Molina, A.; Tantillo, C.; Hughes, S.H.; Arnold, E. Structure and functional implications of the polymerase active site region in a complex of HIV-1 RT with a double-stranded DNA template-primer and an antibody Fab fragment at 2.8 A resolution. J. Mol. Biol. 1998, 284, 1095–1111. [Google Scholar]
  21. Huang, H.; Chopra, R.; Verdine, G.L.; Harrison, S.C. Structure of a covalently trapped catalytic complex of HIV-1 reverse transcriptase: Implications for drug resistance. Science 1998, 282, 1669–1675. [Google Scholar] [CrossRef]
  22. Sarafianos, S.G.; Das, K.; Tantillo, C.; Clark, A.D., Jr.; Ding, J.; Whitcomb, J.M.; Boyer, P.L.; Hughes, S.H.; Arnold, E. Crystal structure of HIV-1 reverse transcriptase in complex with a polypurine tract RNA:DNA. EMBO J. 2001, 20, 1449–1461. [Google Scholar] [CrossRef]
  23. Tuske, S.; Sarafianos, S.G.; Clark, A.D., Jr.; Ding, J.; Naeger, L.K.; White, K.L.; Miller, M.D.; Gibbs, C.S.; Boyer, P.L.; Clark, P.; et al. Structures of HIV-1 RT-DNA complexes before and after incorporation of the anti-AIDS drug tenofovir. Nat. Struct. Mol. Biol. 2004, 11, 469–474. [Google Scholar] [CrossRef]
  24. Das, K.; Bandwar, R.P.; White, K.L.; Feng, J.Y.; Sarafianos, S.G.; Tuske, S.; Tu, X.; Clark, A.D., Jr.; Boyer, P.L.; Hou, X.; et al. Structural basis for the role of the K65R mutation in HIV-1 reverse transcriptase polymerization, excision antagonism, and tenofovir resistance. J. Biol. Chem. 2009, 284, 35092–35100. [Google Scholar] [CrossRef]
  25. Das, K.; Martinez, S.E.; Bauman, J.D.; Arnold, E. HIV-1 reverse transcriptase complex with DNA and nevirapine reveals non-nucleoside inhibition mechanism. Nat. Struct. Mol. Biol. 2012, 19, 253–259. [Google Scholar] [CrossRef]
  26. Jacobo-Molina, A.; Ding, J.; Nanni, R.G.; Clark, A.D., Jr.; Lu, X.; Tantillo, C.; Williams, R.L.; Kamer, G.; Ferris, A.L.; Clark, P.; et al. Crystal structure of human immunodeficiency virus type 1 reverse transcriptase complexed with double-stranded DNA at 3.0 A resolution shows bent DNA. Proc. Natl. Acad. Sci. USA 1993, 90, 6320–6324. [Google Scholar]
  27. Ren, J.; Esnouf, R.; Garman, E.; Somers, D.; Ross, C.; Kirby, I.; Keeling, J.; Darby, G.; Jones, Y.; Stuart, D.; et al. High resolution structures of HIV-1 RT from four RT-inhibitor complexes. Nat. Struct. Biol. 1995, 2, 293–302. [Google Scholar]
  28. Unge, T.; Knight, S.; Bhikhabhai, R.; Lovgren, S.; Dauter, Z.; Wilson, K.; Strandberg, B. 2.2 A resolution structure of the amino-terminal half of HIV-1 reverse transcriptase (fingers and palm subdomains). Structure 1994, 2, 953–961. [Google Scholar] [CrossRef]
  29. Ren, J.; Milton, J.; Weaver, K.L.; Short, S.A.; Stuart, D.I.; Stammers, D.K. Structural basis for the resilience of efavirenz (DMP-266) to drug resistance mutations in HIV-1 reverse transcriptase. Structure 2000, 8, 1089–1094. [Google Scholar] [CrossRef]
  30. Das, K.; Bauman, J.D.; Clark, A.D., Jr.; Frenkel, Y.V.; Lewi, P.J.; Shatkin, A.J.; Hughes, S.H.; Arnold, E. High-resolution structures of HIV-1 reverse transcriptase/TMC278 complexes: Strategic flexibility explains potency against resistance mutations. Proc. Natl. Acad. Sci. USA 2008, 105, 1466–1471. [Google Scholar] [CrossRef]
  31. Pata, J.D.; Stirtan, W.G.; Goldstein, S.W.; Steitz, T.A. Structure of HIV-1 reverse transcriptase bound to an inhibitor active against mutant reverse transcriptases resistant to other nonnucleoside inhibitors. Proc. Natl. Acad. Sci. USA 2004, 101, 10548–10553. [Google Scholar] [CrossRef]
  32. Hsiou, Y.; Ding, J.; Das, K.; Clark, A.D., Jr.; Hughes, S.H.; Arnold, E. Structure of unliganded HIV-1 reverse transcriptase at 2.7 A resolution: Implications of conformational changes for polymerization and inhibition mechanisms. Structure 1996, 4, 853–860. [Google Scholar]
  33. Kohlstaedt, L.A.; Wang, J.; Friedman, J.M.; Rice, P.A.; Steitz, T.A. Crystal structure at 3.5 A resolution of HIV-1 reverse transcriptase complexed with an inhibitor. Science 1992, 256, 1783–1790. [Google Scholar]
  34. Lapkouski, M.; Tian, L.; Miller, J.T.; Le Grice, S.F.; Yang, W. Complexes of HIV-1 RT, NNRTI and RNA/DNA hybrid reveal a structure compatible with RNA degradation. Nat. Struct. Mol. Biol. 2013, 20, 230–236. [Google Scholar] [CrossRef]
  35. Das, K.; Sarafianos, S.G.; Arnold, E. Structural requirements for RNA degradation by HIV-1 reverse transcriptase. Nat. Struct. Mol. Biol. 2013, 20, 1341–1342. [Google Scholar] [CrossRef]
  36. The PyMOL Molecular Graphics System, version 1.5.0.4; Schrödinger, LLC: New York, NY, USA, 2014.
  37. Ren, J.; Bird, L.E.; Chamberlain, P.P.; Stewart-Jones, G.B.; Stuart, D.I.; Stammers, D.K. Structure of HIV-2 reverse transcriptase at 2.35-A resolution and the mechanism of resistance to non-nucleoside inhibitors. Proc. Natl. Acad. Sci. USA 2002, 99, 14410–14415. [Google Scholar]
  38. Schuckmann, M.M.; Marchand, B.; Hachiya, A.; Kodama, E.N.; Kirby, K.A.; Singh, K.; Sarafianos, S.G. The N348I mutation at the connection subdomain of HIV-1 reverse transcriptase decreases binding to nevirapine. J. Biol. Chem. 2010, 285, 38700–38709. [Google Scholar] [CrossRef]
  39. Chung, S.; Miller, J.T.; Lapkouski, M.; Tian, L.; Yang, W.; Le Grice, S.F. Examining the role of the HIV-1 reverse transcriptase p51 subunit in positioning and hydrolysis of RNA/DNA hybrids. J. Biol. Chem. 2013, 288, 16177–16184. [Google Scholar] [CrossRef]
  40. Singh, K.; Marchand, B.; Kirby, K.A.; Michailidis, E.; Sarafianos, S.G. Structural aspects of drug resistance and inhibition of HIV-1 Reverse Transcriptase. Viruses 2010, 2, 606–638. [Google Scholar] [CrossRef]
  41. De Clercq, E. Antiviral drugs in current clinical use. J. Clin. Virol. 2004, 30, 115–133. [Google Scholar]
  42. Parniak, M.A.; Sluis-Cremer, N. Inhibitors of HIV-1 reverse transcriptase. Adv. Pharmacol. 2000, 49, 67–109. [Google Scholar]
  43. Sarafianos, S.G.; Marchand, B.; Das, K.; Himmel, D.M.; Parniak, M.A.; Hughes, S.H.; Arnold, E. Structure and function of HIV-1 reverse transcriptase: Molecular mechanisms of polymerization and inhibition. J. Mol. Biol. 2009, 385, 693–713. [Google Scholar] [CrossRef]
  44. Sarafianos, S.G.; Das, K.; Clark, A.D., Jr.; Ding, J.; Boyer, P.L.; Hughes, S.H.; Arnold, E. Lamivudine (3TC) resistance in HIV-1 reverse transcriptase involves steric hindrance with beta-branched amino acids. Proc. Natl. Acad. Sci. USA 1999, 96, 10027–10032. [Google Scholar] [CrossRef]
  45. Smerdon, S.J.; Jager, J.; Wang, J.; Kohlstaedt, L.A.; Chirino, A.J.; Friedman, J.M.; Rice, P.A.; Steitz, T.A. Structure of the binding site for nonnucleoside inhibitors of the reverse transcriptase of human immunodeficiency virus type 1. Proc. Natl. Acad. Sci. USA 1994, 91, 3911–3915. [Google Scholar] [CrossRef]
  46. Tantillo, C.; Ding, J.; Jacobo-Molina, A.; Nanni, R.G.; Boyer, P.L.; Hughes, S.H.; Pauwels, R.; Andries, K.; Janssen, P.A.; Arnold, E. Locations of anti-AIDS drug binding sites and resistance mutations in the three-dimensional structure of HIV-1 reverse transcriptase. Implications for mechanisms of drug inhibition and resistance. J. Mol. Biol. 1994, 243, 369–387. [Google Scholar]
  47. Johnson, V.A.; Calvez, V.; Gunthard, H.F.; Paredes, R.; Pillay, D.; Shafer, R.W.; Wensing, A.M.; Richman, D.D. Update of the drug resistance mutations in HIV-1: March 2013. Top. Antivir. Med. 2013, 21, 6–14. [Google Scholar]
  48. Miller, M.D.; Margot, N.; Lu, B.; Zhong, L.; Chen, S.S.; Cheng, A.; Wulfsohn, M. Genotypic and phenotypic predictors of the magnitude of response to tenofovir disoproxil fumarate treatment in antiretroviral-experienced patients. J. Infect. Dis. 2004, 189, 837–846. [Google Scholar] [CrossRef]
  49. Whitcomb, J.M.; Parkin, N.T.; Chappey, C.; Hellmann, N.S.; Petropoulos, C.J. Broad nucleoside reverse-transcriptase inhibitor cross-resistance in human immunodeficiency virus type 1 clinical isolates. J. Infect. Dis. 2003, 188, 992–1000. [Google Scholar] [CrossRef]
  50. Kuritzkes, D.R.; Bassett, R.L.; Hazelwood, J.D.; Barrett, H.; Rhodes, R.A.; Young, R.K.; Johnson, V.A.; Adult, A.P.T. Rate of thymidine analogue resistance mutation accumulation with zidovudine- or stavudine-based regimens. J. Acquir. Immune Defic. Syndr. 2004, 36, 600–603. [Google Scholar] [CrossRef]
  51. ACD/ChemSketch Advanced Chemistry Development, I.: Toronto, ON, Canada. Available online: http://www.acdlabs.com/ (accessed on 7 July 2014).
  52. Deval, J.; Selmi, B.; Boretto, J.; Egloff, M.P.; Guerreiro, C.; Sarfati, S.; Canard, B. The molecular mechanism of multidrug resistance by the Q151M human immunodeficiency virus type 1 reverse transcriptase and its suppression using alpha-boranophosphate nucleotide analogues. J. Biol. Chem. 2002, 277, 42097–42104. [Google Scholar] [CrossRef]
  53. Deval, J.; Navarro, J.M.; Selmi, B.; Courcambeck, J.; Boretto, J.; Halfon, P.; Garrido-Urbani, S.; Sire, J.; Canard, B. A loss of viral replicative capacity correlates with altered DNA polymerization kinetics by the human immunodeficiency virus reverse transcriptase bearing the K65R and L74V dideoxynucleoside resistance substitutions. J. Biol. Chem. 2004, 279, 25489–25496. [Google Scholar] [CrossRef]
  54. Deval, J.; White, K.L.; Miller, M.D.; Parkin, N.T.; Courcambeck, J.; Halfon, P.; Selmi, B.; Boretto, J.; Canard, B. Mechanistic basis for reduced viral and enzymatic fitness of HIV-1 reverse transcriptase containing both K65R and M184V mutations. J. Biol. Chem. 2004, 279, 509–516. [Google Scholar] [CrossRef]
  55. Sluis-Cremer, N.; Sheen, C.W.; Zelina, S.; Torres, P.S.; Parikh, U.M.; Mellors, J.W. Molecular mechanism by which the K70E mutation in human immunodeficiency virus type 1 reverse transcriptase confers resistance to nucleoside reverse transcriptase inhibitors. Antimicrob. Agents Chemother. 2007, 51, 48–53. [Google Scholar] [CrossRef]
  56. Das, K.; Arnold, E. HIV-1 reverse transcriptase and antiviral drug resistance. Part 2. Curr. Opin. Virol. 2013, 3, 119–128. [Google Scholar] [CrossRef]
  57. Krebs, R.; Immendorfer, U.; Thrall, S.H.; Wohrl, B.M.; Goody, R.S. Single-step kinetics of HIV-1 reverse transcriptase mutants responsible for virus resistance to nucleoside inhibitors zidovudine and 3-TC. Biochemistry 1997, 36, 10292–10300. [Google Scholar] [CrossRef]
  58. St Clair, M.H.; Martin, J.L.; Tudor-Williams, G.; Bach, M.C.; Vavro, C.L.; King, D.M.; Kellam, P.; Kemp, S.D.; Larder, B.A. Resistance to ddI and sensitivity to AZT induced by a mutation in HIV-1 reverse transcriptase. Science 1991, 253, 1557–1559. [Google Scholar] [CrossRef]
  59. Martin, J.L.; Wilson, J.E.; Haynes, R.L.; Furman, P.A. Mechanism of resistance of human immunodeficiency virus type 1 to 2',3'-dideoxyinosine. Proc. Natl. Acad. Sci. USA 1993, 90, 6135–6139. [Google Scholar] [CrossRef]
  60. Matamoros, T.; Kim, B.; Menendez-Arias, L. Mechanistic insights into the role of Val75 of HIV-1 reverse transcriptase in misinsertion and mispair extension fidelity of DNA synthesis. J. Mol. Biol. 2008, 375, 1234–1248. [Google Scholar] [CrossRef]
  61. Ueno, T.; Mitsuya, H. Comparative enzymatic study of HIV-1 reverse transcriptase resistant to 2',3'-dideoxynucleotide analogs using the single-nucleotide incorporation assay. Biochemistry 1997, 36, 1092–1099. [Google Scholar] [CrossRef]
  62. Tisdale, M.; Kemp, S.D.; Parry, N.R.; Larder, B.A. Rapid in vitro selection of human immunodeficiency virus type 1 resistant to 3'-thiacytidine inhibitors due to a mutation in the YMDD region of reverse transcriptase. Proc. Natl. Acad. Sci. USA 1993, 90, 5653–5656. [Google Scholar] [CrossRef]
  63. Selmi, B.; Boretto, J.; Sarfati, S.R.; Guerreiro, C.; Canard, B. Mechanism-based suppression of dideoxynucleotide resistance by K65R human immunodeficiency virus reverse transcriptase using an alpha-boranophosphate nucleoside analogue. J. Biol. Chem. 2001, 276, 48466–48472. [Google Scholar]
  64. Feng, J.Y.; Myrick, F.T.; Margot, N.A.; Mulamba, G.B.; Rimsky, L.; Borroto-Esoda, K.; Selmi, B.; Canard, B. Virologic and enzymatic studies revealing the mechanism of K65R- and Q151M-associated HIV-1 drug resistance towards emtricitabine and lamivudine. Nucleosides Nucleotides Nucleic Acids 2006, 25, 89–107. [Google Scholar] [CrossRef]
  65. Shirasaka, T.; Kavlick, M.F.; Ueno, T.; Gao, W.Y.; Kojima, E.; Alcaide, M.L.; Chokekijchai, S.; Roy, B.M.; Arnold, E.; Yarchoan, R.; et al. Emergence of human immunodeficiency virus type 1 variants with resistance to multiple dideoxynucleosides in patients receiving therapy with dideoxynucleosides. Proc. Natl. Acad. Sci. USA 1995, 92, 2398–2402. [Google Scholar] [CrossRef]
  66. Tchesnokov, E.P.; Obikhod, A.; Massud, I.; Lisco, A.; Vanpouille, C.; Brichacek, B.; Balzarini, J.; McGuigan, C.; Derudas, M.; Margolis, L.; et al. Mechanisms associated with HIV-1 resistance to acyclovir by the V75I mutation in reverse transcriptase. J. Biol. Chem. 2009, 284, 21496–21504. [Google Scholar] [CrossRef]
  67. Frangeul, A.; Bussetta, C.; Deval, J.; Barral, K.; Alvarez, K.; Canard, B. Gln151 of HIV-1 reverse transcriptase acts as a steric gate towards clinically relevant acyclic phosphonate nucleotide analogues. Antivir. Ther. 2008, 13, 115–124. [Google Scholar]
  68. Meyer, P.R.; Matsuura, S.E.; So, A.G.; Scott, W.A. Unblocking of chain-terminated primer by HIV-1 reverse transcriptase through a nucleotide-dependent mechanism. Proc. Natl. Acad. Sci. USA 1998, 95, 13471–13476. [Google Scholar] [CrossRef]
  69. Arion, D.; Kaushik, N.; McCormick, S.; Borkow, G.; Parniak, M.A. Phenotypic mechanism of HIV-1 resistance to 3'-azido-3'-deoxythymidine (AZT): Increased polymerization processivity and enhanced sensitivity to pyrophosphate of the mutant viral reverse transcriptase. Biochemistry 1998, 37, 15908–15917. [Google Scholar] [CrossRef]
  70. Sarafianos, S.G.; Clark, A.D., Jr.; Das, K.; Tuske, S.; Birktoft, J.J.; Ilankumaran, P.; Ramesha, A.R.; Sayer, J.M.; Jerina, D.M.; Boyer, P.L.; et al. Structures of HIV-1 reverse transcriptase with pre- and post-translocation AZTMP-terminated DNA. EMBO J. 2002, 21, 6614–6624. [Google Scholar] [CrossRef]
  71. Sarafianos, S.G.; Clark, A.D., Jr.; Tuske, S.; Squire, C.J.; Das, K.; Sheng, D.; Ilankumaran, P.; Ramesha, A.R.; Kroth, H.; Sayer, J.M.; et al. Trapping HIV-1 reverse transcriptase before and after translocation on DNA. J. Biol. Chem. 2003, 278, 16280–16288. [Google Scholar] [CrossRef]
  72. Das, K.; Arnold, E. HIV-1 reverse transcriptase and antiviral drug resistance. Part 1. Curr. Opin. Virol. 2013, 3, 111–118. [Google Scholar] [CrossRef]
  73. Lacey, S.F.; Reardon, J.E.; Furfine, E.S.; Kunkel, T.A.; Bebenek, K.; Eckert, K.A.; Kemp, S.D.; Larder, B.A. Biochemical studies on the reverse transcriptase and RNase H activities from human immunodeficiency virus strains resistant to 3'-azido-3'-deoxythymidine. J. Biol. Chem. 1992, 267, 15789–15794. [Google Scholar]
  74. Betancor, G.; Nevot, M.; Mendieta, J.; Gomez-Puertas, P.; Martinez, M.A.; Menendez-Arias, L. Molecular basis of the association of H208Y and thymidine analogue resistance mutations M41L, L210W and T215Y in the HIV-1 reverse transcriptase of treated patients. Antivir. Res. 2014, 106, 42–52. [Google Scholar] [CrossRef]
  75. Arion, D.; Parniak, M.A. HIV resistance to zidovudine: The role of pyrophosphorolysis. Drug Resist. Updat 1999, 2, 91–95. [Google Scholar] [CrossRef]
  76. Mas, A.; Parera, M.; Briones, C.; Soriano, V.; Martinez, M.A.; Domingo, E.; Menendez-Arias, L. Role of a dipeptide insertion between codons 69 and 70 of HIV-1 reverse transcriptase in the mechanism of AZT resistance. EMBO J. 2000, 19, 5752–5761. [Google Scholar] [CrossRef]
  77. Boyer, P.L.; Imamichi, T.; Sarafianos, S.G.; Arnold, E.; Hughes, S.H. Effects of the Delta67 complex of mutations in human immunodeficiency virus type 1 reverse transcriptase on nucleoside analog excision. J. Virol. 2004, 78, 9987–9997. [Google Scholar] [CrossRef]
  78. Meyer, P.R.; Matsuura, S.E.; Mian, A.M.; So, A.G.; Scott, W.A. A mechanism of AZT resistance: An increase in nucleotide-dependent primer unblocking by mutant HIV-1 reverse transcriptase. Mol. Cell 1999, 4, 35–43. [Google Scholar] [CrossRef]
  79. Meyer, P.R.; Matsuura, S.E.; Schinazi, R.F.; So, A.G.; Scott, W.A. Differential removal of thymidine nucleotide analogues from blocked DNA chains by human immunodeficiency virus reverse transcriptase in the presence of physiological concentrations of 2'-deoxynucleoside triphosphates. Antimicrob. Agents Chemother. 2000, 44, 3465–3472. [Google Scholar] [CrossRef]
  80. Lai, M.T.; Lu, M.; Felock, P.J.; Hrin, R.C.; Wang, Y.J.; Yan, Y.; Munshi, S.; McGaughey, G.B.; Tynebor, R.M.; Tucker, T.J.; et al. Distinct mutation pathways of non-subtype B HIV-1 during in vitro resistance selection with nonnucleoside reverse transcriptase inhibitors. Antimicrob. Agents Chemother. 2010, 54, 4812–4824. [Google Scholar] [CrossRef]
  81. Boyer, P.L.; Sarafianos, S.G.; Arnold, E.; Hughes, S.H. Selective excision of AZTMP by drug-resistant human immunodeficiency virus reverse transcriptase. J. Virol. 2001, 75, 4832–4842. [Google Scholar] [CrossRef]
  82. Arion, D.; Sluis-Cremer, N.; Parniak, M.A. Mechanism by which phosphonoformic acid resistance mutations restore 3'-azido-3'-deoxythymidine (AZT) sensitivity to AZT-resistant HIV-1 reverse transcriptase. J. Biol. Chem. 2000, 275, 9251–9255. [Google Scholar] [CrossRef]
  83. Sarafianos, S.G.; Hughes, S.H.; Arnold, E. Designing anti-AIDS drugs targeting the major mechanism of HIV-1 RT resistance to nucleoside analog drugs. Int. J. Biochem. Cell Biol. 2004, 36, 1706–1715. [Google Scholar] [CrossRef]
  84. Novitsky, V.; Wester, C.W.; DeGruttola, V.; Bussmann, H.; Gaseitsiwe, S.; Thomas, A.; Moyo, S.; Musonda, R.; van Widenfelt, E.; Marlink, R.G.; et al. The reverse transcriptase 67N 70R 215Y genotype is the predominant TAM pathway associated with virologic failure among HIV type 1C-infected adults treated with ZDV/ddI-containing HAART in southern Africa. AIDS Res. Hum. Retroviruses 2007, 23, 868–878. [Google Scholar] [CrossRef]
  85. Deshpande, A.; Jauvin, V.; Magnin, N.; Pinson, P.; Faure, M.; Masquelier, B.; Aurillac-Lavignolle, V.; Fleury, H.J. Resistance mutations in subtype C HIV type 1 isolates from Indian patients of Mumbai receiving NRTIs plus NNRTIs and experiencing a treatment failure: Resistance to AR. AIDS Res. Hum. Retroviruses 2007, 23, 335–340. [Google Scholar] [CrossRef]
  86. Hosseinipour, M.C.; van Oosterhout, J.J.; Weigel, R.; Phiri, S.; Kamwendo, D.; Parkin, N.; Fiscus, S.A.; Nelson, J.A.; Eron, J.J.; Kumwenda, J. The public health approach to identify antiretroviral therapy failure: High-level nucleoside reverse transcriptase inhibitor resistance among Malawians failing first-line antiretroviral therapy. AIDS 2009, 23, 1127–1134. [Google Scholar]
  87. Barth, R.E.; Wensing, A.M.; Tempelman, H.A.; Moraba, R.; Schuurman, R.; Hoepelman, A.I. Rapid accumulation of nonnucleoside reverse transcriptase inhibitor-associated resistance: Evidence of transmitted resistance in rural South Africa. AIDS 2008, 22, 2210–2212. [Google Scholar] [CrossRef]
  88. Doualla-Bell, F.; Avalos, A.; Brenner, B.; Gaolathe, T.; Mine, M.; Gaseitsiwe, S.; Oliveira, M.; Moisi, D.; Ndwapi, N.; Moffat, H.; et al. High prevalence of the K65R mutation in human immunodeficiency virus type 1 subtype C isolates from infected patients in Botswana treated with didanosine-based regimens. Antimicrob. Agents Chemother. 2006, 50, 4182–4185. [Google Scholar] [CrossRef]
  89. Sunpath, H.; Wu, B.; Gordon, M.; Hampton, J.; Johnson, B.; Moosa, M.Y.; Ordonez, C.; Kuritzkes, D.R.; Marconi, V.C. High rate of K65R for antiretroviral therapy-naive patients with subtype C HIV infection failing a tenofovir-containing first-line regimen. AIDS 2012, 26, 1679–1684. [Google Scholar] [CrossRef]
  90. Brenner, B.G.; Coutsinos, D. The K65R mutation in HIV-1 reverse transcriptase: Genetic barriers, resistance profile and clinical implications. HIV Ther. 2009, 3, 583–594. [Google Scholar] [CrossRef]
  91. Orrell, C.; Walensky, R.P.; Losina, E.; Pitt, J.; Freedberg, K.A.; Wood, R. HIV type-1 clade C resistance genotypes in treatment-naive patients and after first virological failure in a large community antiretroviral therapy programme. Antivir. Ther. 2009, 14, 523–531. [Google Scholar]
  92. Turner, D.; Shahar, E.; Katchman, E.; Kedem, E.; Matus, N.; Katzir, M.; Hassoun, G.; Pollack, S.; Kessner, R.; Wainberg, M.A.; et al. Prevalence of the K65R resistance reverse transcriptase mutation in different HIV-1 subtypes in Israel. J. Med. Virol. 2009, 81, 1509–1512. [Google Scholar] [CrossRef]
  93. Deshpande, A.; Jeannot, A.C.; Schrive, M.H.; Wittkop, L.; Pinson, P.; Fleury, H.J. Analysis of RT sequences of subtype C HIV-type 1 isolates from indian patients at failure of a first-line treatment according to clinical and/or immunological WHO guidelines. AIDS Res. Hum. Retroviruses 2010, 26, 343–350. [Google Scholar] [CrossRef]
  94. Gupta, R.K.; Chrystie, I.L.; O’Shea, S.; Mullen, J.E.; Kulasegaram, R.; Tong, C.Y. K65R and Y181C are less prevalent in HAART-experienced HIV-1 subtype A patients. AIDS 2005, 19, 1916–1919. [Google Scholar] [CrossRef]
  95. Fontella, R.; Soares, M.A.; Schrago, C.G. On the origin of HIV-1 subtype C in South America. AIDS 2008, 22, 2001–2011. [Google Scholar] [CrossRef]
  96. Coutsinos, D.; Invernizzi, C.F.; Moisi, D.; Oliveira, M.; Martinez-Cajas, J.L.; Brenner, B.G.; Wainberg, M.A. A template-dependent dislocation mechanism potentiates K65R reverse transcriptase mutation development in subtype C variants of HIV-1. PLoS One 2011, 6, e20208. [Google Scholar] [CrossRef]
  97. Coutsinos, D.; Invernizzi, C.F.; Xu, H.; Moisi, D.; Oliveira, M.; Brenner, B.G.; Wainberg, M.A. Template usage is responsible for the preferential acquisition of the K65R reverse transcriptase mutation in subtype C variants of human immunodeficiency virus type 1. J. Virol. 2009, 83, 2029–2033. [Google Scholar] [CrossRef]
  98. Antinori, A.; Zaccarelli, M.; Cingolani, A.; Forbici, F.; Rizzo, M.G.; Trotta, M.P.; di Giambenedetto, S.; Narciso, P.; Ammassari, A.; Girardi, E.; et al. Cross-resistance among nonnucleoside reverse transcriptase inhibitors limits recycling efavirenz after nevirapine failure. AIDS Res. Hum. Retroviruses 2002, 18, 835–838. [Google Scholar] [CrossRef]
  99. Brenner, B.; Turner, D.; Oliveira, M.; Moisi, D.; Detorio, M.; Carobene, M.; Marlink, R.G.; Schapiro, J.; Roger, M.; Wainberg, M.A. A V106M mutation in HIV-1 clade C viruses exposed to efavirenz confers cross-resistance to non-nucleoside reverse transcriptase inhibitors. AIDS 2003, 17, F1–F5. [Google Scholar] [CrossRef]
  100. Lazzarin, A.; Campbell, T.; Clotet, B.; Johnson, M.; Katlama, C.; Moll, A.; Towner, W.; Trottier, B.; Peeters, M.; Vingerhoets, J.; et al. Efficacy and safety of TMC125 (etravirine) in treatment-experienced HIV-1-infected patients in DUET-2: 24-week results from a randomised, double-blind, placebo-controlled trial. Lancet 2007, 370, 39–48. [Google Scholar] [CrossRef]
  101. Madruga, J.V.; Cahn, P.; Grinsztejn, B.; Haubrich, R.; Lalezari, J.; Mills, A.; Pialoux, G.; Wilkin, T.; Peeters, M.; Vingerhoets, J.; et al. Efficacy and safety of TMC125 (etravirine) in treatment-experienced HIV-1-infected patients in DUET-1: 24-week results from a randomised, double-blind, placebo-controlled trial. Lancet 2007, 370, 29–38. [Google Scholar] [CrossRef]
  102. Janssen, P.A.; Lewi, P.J.; Arnold, E.; Daeyaert, F.; de Jonge, M.; Heeres, J.; Koymans, L.; Vinkers, M.; Guillemont, J.; Pasquier, E.; et al. In search of a novel anti-HIV drug: Multidisciplinary coordination in the discovery of 4-[[4-[[4-[(1E)-2-cyanoethenyl]-2,6-dimethylphenyl]amino]-2- pyrimidinyl]amino]benzonitrile (R278474, rilpivirine). J. Med. Chem. 2005, 48, 1901–1909. [Google Scholar] [CrossRef]
  103. Michailidis, E.; Singh, K.; Ryan, E.M.; Hachiya, A.; Ong, Y.T.; Kirby, K.A.; Marchand, B.; Kodama, E.N.; Mitsuya, H.; Parniak, M.A.; et al. Effect of translocation defective reverse transcriptase inhibitors on the activity of N348I, a connection subdomain drug resistant HIV-1 reverse transcriptase mutant. Cell Mol. Biol. (Noisy-le-grand) 2012, 58, 187–195. [Google Scholar]
  104. Kulkarni, R.; Babaoglu, K.; Lansdon, E.B.; Rimsky, L.; Van Eygen, V.; Picchio, G.; Svarovskaia, E.; Miller, M.D.; White, K.L. The HIV-1 reverse transcriptase M184I mutation enhances the E138K-associated resistance to rilpivirine and decreases viral fitness. J. Acquir. Immune Defic. Syndr. 2012, 59, 47–54. [Google Scholar] [CrossRef]
  105. Spence, R.A.; Kati, W.M.; Anderson, K.S.; Johnson, K.A. Mechanism of inhibition of HIV-1 reverse transcriptase by nonnucleoside inhibitors. Science 1995, 267, 988–993. [Google Scholar] [CrossRef]
  106. Spence, R.A.; Anderson, K.S.; Johnson, K.A. HIV-1 reverse transcriptase resistance to nonnucleoside inhibitors. Biochemistry 1996, 35, 1054–1063. [Google Scholar] [CrossRef]
  107. Maga, G.; Amacker, M.; Ruel, N.; Hubscher, U.; Spadari, S. Resistance to nevirapine of HIV-1 reverse transcriptase mutants: Loss of stabilizing interactions and thermodynamic or steric barriers are induced by different single amino acid substitutions. J. Mol. Biol. 1997, 274, 738–747. [Google Scholar] [CrossRef]
  108. Ren, J.; Nichols, C.; Bird, L.; Chamberlain, P.; Weaver, K.; Short, S.; Stuart, D.I.; Stammers, D.K. Structural mechanisms of drug resistance for mutations at codons 181 and 188 in HIV-1 reverse transcriptase and the improved resilience of second generation non-nucleoside inhibitors. J. Mol. Biol. 2001, 312, 795–805. [Google Scholar]
  109. Ren, J.; Stammers, D.K. Structural basis for drug resistance mechanisms for non-nucleoside inhibitors of HIV reverse transcriptase. Virus Res. 2008, 134, 157–170. [Google Scholar] [CrossRef]
  110. Brenner, B.G.; Oliveira, M.; Doualla-Bell, F.; Moisi, D.D.; Ntemgwa, M.; Frankel, F.; Essex, M.; Wainberg, M.A. HIV-1 subtype C viruses rapidly develop K65R resistance to tenofovir in cell culture. AIDS 2006, 20, F9–F13. [Google Scholar] [CrossRef]
  111. Arrive, E.; Newell, M.L.; Ekouevi, D.K.; Chaix, M.L.; Thiebaut, R.; Masquelier, B.; Leroy, V.; Perre, P.V.; Rouzioux, C.; Dabis, F.; et al. Prevalence of resistance to nevirapine in mothers and children after single-dose exposure to prevent vertical transmission of HIV-1: A meta-analysis. Int. J. Epidemiol. 2007, 36, 1009–1021. [Google Scholar] [CrossRef]
  112. Eshleman, S.H.; Guay, L.A.; Wang, J.; Mwatha, A.; Brown, E.R.; Musoke, P.; Mmiro, F.; Jackson, J.B. Distinct patterns of emergence and fading of K103N and Y181C in women with subtype A vs. D after single-dose nevirapine: HIVNET 012. J. Acquir. Immune Defic. Syndr. 2005; 40, 24–29. [Google Scholar]
  113. Eshleman, S.H.; Mracna, M.; Guay, L.A.; Deseyve, M.; Cunningham, S.; Mirochnick, M.; Musoke, P.; Fleming, T.; Glenn Fowler, M.; Mofenson, L.M.; et al. Selection and fading of resistance mutations in women and infants receiving nevirapine to prevent HIV-1 vertical transmission (HIVNET 012). AIDS 2001, 15, 1951–1957. [Google Scholar] [CrossRef]
  114. Loemba, H.; Brenner, B.; Parniak, M.A.; Ma'ayan, S.; Spira, B.; Moisi, D.; Oliveira, M.; Detorio, M.; Wainberg, M.A. Genetic divergence of human immunodeficiency virus type 1 Ethiopian clade C reverse transcriptase (RT) and rapid development of resistance against nonnucleoside inhibitors of RT. Antimicrob. Agents Chemother. 2002, 46, 2087–2094. [Google Scholar] [CrossRef]
  115. Marconi, V.C.; Sunpath, H.; Lu, Z.; Gordon, M.; Koranteng-Apeagyei, K.; Hampton, J.; Carpenter, S.; Giddy, J.; Ross, D.; Holst, H.; et al. Prevalence of HIV-1 drug resistance after failure of a first highly active antiretroviral therapy regimen in KwaZulu Natal, South Africa. Clin. Infect. Dis. 2008, 46, 1589–1597. [Google Scholar] [CrossRef]
  116. Ly, N.; Recordon-Pinson, P.; Phoung, V.; Srey, C.; Kruy, L.S.; Koum, K.; Chhum, V.; Glaziou, P.; Fleury, H.J.; Reynes, J.M. Characterization of mutations in HIV type 1 isolates from 144 Cambodian recently infected patients and pregnant women naive to antiretroviral drugs. AIDS Res. Hum. Retroviruses 2005, 21, 971–976. [Google Scholar] [CrossRef]
  117. Rajesh, L.; Karunaianantham, R.; Narayanan, P.R.; Swaminathan, S. Antiretroviral drug-resistant mutations at baseline and at time of failure of antiretroviral therapy in HIV type 1-coinfected TB patients. AIDS Res. Hum. Retroviruses 2009, 25, 1179–1185. [Google Scholar] [CrossRef]
  118. Grossman, Z.; Istomin, V.; Averbuch, D.; Lorber, M.; Risenberg, K.; Levi, I.; Chowers, M.; Burke, M.; Bar Yaacov, N.; Schapiro, J.M. Genetic variation at NNRTI resistance-associated positions in patients infected with HIV-1 subtype C. AIDS 2004, 18, 909–915. [Google Scholar] [CrossRef]
  119. Cohen, C.J.; Andrade-Villanueva, J.; Clotet, B.; Fourie, J.; Johnson, M.A.; Ruxrungtham, K.; Wu, H.; Zorrilla, C.; Crauwels, H.; Rimsky, L.T.; et al. Rilpivirine versus efavirenz with two background nucleoside or nucleotide reverse transcriptase inhibitors in treatment-naive adults infected with HIV-1 (THRIVE): A phase 3, randomised, non-inferiority trial. Lancet 2011, 378, 229–237. [Google Scholar] [CrossRef]
  120. Hu, Z.; Kuritzkes, D.R. Interaction of reverse transcriptase (RT) mutations conferring resistance to lamivudine and etravirine: Effects on fitness and RT activity of human immunodeficiency virus type 1. J. Virol. 2011, 85, 11309–11314. [Google Scholar] [CrossRef]
  121. Xu, H.T.; Asahchop, E.L.; Oliveira, M.; Quashie, P.K.; Quan, Y.; Brenner, B.G.; Wainberg, M.A. Compensation by the E138K mutation in HIV-1 reverse transcriptase for deficits in viral replication capacity and enzyme processivity associated with the M184I/V mutations. J. Virol. 2011, 85, 11300–11308. [Google Scholar] [CrossRef]
  122. Maiga, A.I.; Descamps, D.; Morand-Joubert, L.; Malet, I.; Derache, A.; Cisse, M.; Koita, V.; Akonde, A.; Diarra, B.; Wirden, M.; et al. Resistance-associated mutations to etravirine (TMC-125) in antiretroviral-naive patients infected with non-B HIV-1 subtypes. Antimicrob. Agents Chemother. 2010, 54, 728–733. [Google Scholar] [CrossRef]
  123. Asahchop, E.L.; Oliveira, M.; Wainberg, M.A.; Brenner, B.G.; Moisi, D.; Toni, T.; Tremblay, C.L. Characterization of the E138K resistance mutation in HIV-1 reverse transcriptase conferring susceptibility to etravirine in B and non-B HIV-1 subtypes. Antimicrob. Agents Chemother. 2011, 55, 600–607. [Google Scholar] [CrossRef]
  124. Neogi, U.; Shet, A.; Shamsundar, R.; Ekstrand, M.L. Selection of nonnucleoside reverse transcriptase inhibitor-associated mutations in HIV-1 subtype C: Evidence of etravirine cross-resistance. AIDS 2011, 25, 1123–1126. [Google Scholar] [CrossRef]
  125. Lambert-Niclot, S.; Charpentier, C.; Storto, A.; Fofana, D.; Soulie, C.; Fourati, S.; Wirden, M.; Morand-Joubert, L.; Masquelier, B.; Flandre, P.; et al. Rilpivirine, emtricitabine and tenofovir resistance in HIV-1-infected rilpivirine-naive patients failing antiretroviral therapy. J. Antimicrob. Chemother. 2014, 69, 1086–1089. [Google Scholar] [CrossRef]
  126. Christen, M.T.; Menon, L.; Myshakina, N.S.; Ahn, J.; Parniak, M.A.; Ishima, R. Structural basis of the allosteric inhibitor interaction on the HIV-1 reverse transcriptase RNase H domain. Chem. Biol. Drug Des. 2012, 80, 706–716, PMCID: PMC3465473. [Google Scholar]
  127. European Medicines Agency. Edurant® 25 mg Film-coated Tablets: Summary of Product Characteristics. Available online: http://www.ema.europa.eu/docs/en_GB/document_library/EPAR_-_Product_Information/human/002264/WC500118874.pdf (accessed on 5 May 2014).
  128. Wensing, A.M.; van de Vijver, D.A.; Angarano, G.; Asjo, B.; Balotta, C.; Boeri, E.; Camacho, R.; Chaix, M.L.; Costagliola, D.; De Luca, A.; et al. Prevalence of drug-resistant HIV-1 variants in untreated individuals in Europe: Implications for clinical management. J. Infect. Dis. 2005, 192, 958–966. [Google Scholar] [CrossRef]
  129. Sluis-Cremer, N.; Jordan, M.R.; Huber, K.; Wallis, C.L.; Bertagnolio, S.; Mellors, J.W.; Parkin, N.T.; Harrigan, P.R. E138A in HIV-1 reverse transcriptase is more common in subtype C than B: Implications for rilpivirine use in resource-limited settings. Antivir. Res. 2014, 107, 31–34. [Google Scholar] [CrossRef]
  130. Bunupuradah, T.; Ananworanich, J.; Chetchotisakd, P.; Kantipong, P.; Jirajariyavej, S.; Sirivichayakul, S.; Munsakul, W.; Prasithsirikul, W.; Sungkanuparph, S.; Bowonwattanuwong, C.; et al. Etravirine and rilpivirine resistance in HIV-1 subtype CRF01_AE-infected adults failing non-nucleoside reverse transcriptase inhibitor-based regimens. Antivir. Ther. 2011, 16, 1113–1121. [Google Scholar] [CrossRef]
  131. Kantor, R.; Katzenstein, D.A.; Efron, B.; Carvalho, A.P.; Wynhoven, B.; Cane, P.; Clarke, J.; Sirivichayakul, S.; Soares, M.A.; Snoeck, J.; et al. Impact of HIV-1 subtype and antiretroviral therapy on protease and reverse transcriptase genotype: Results of a global collaboration. PLoS Med 2005, 2, e112. [Google Scholar]
  132. Santoro, M.M.; Perno, C.F. HIV-1 Genetic Variability and Clinical Implications. ISRN Microbiol 2013, 2013, 481314. [Google Scholar] [CrossRef]
  133. Santoro, M.M.; Alteri, C.; Ronga, L.; Flandre, P.; Fabeni, L.; Mercurio, F.; D'Arrigo, R.; Gori, C.; Palamara, G.; Bertoli, A.; et al. Comparative analysis of drug resistance among B and the most prevalent non-B HIV type 1 subtypes (C, F, and CRF02_AG) in Italy. AIDS Res. Hum. Retroviruses 2012, 28, 1285–1293. [Google Scholar] [CrossRef]
  134. Flys, T.S.; Chen, S.; Jones, D.C.; Hoover, D.R.; Church, J.D.; Fiscus, S.A.; Mwatha, A.; Guay, L.A.; Mmiro, F.; Musoke, P.; et al. Quantitative analysis of HIV-1 variants with the K103N resistance mutation after single-dose nevirapine in women with HIV-1 subtypes A, C, and D. J. Acquir. Immune Defic. Syndr. 2006, 42, 610–613. [Google Scholar] [CrossRef]
  135. Johnson, J.A.; Li, J.F.; Morris, L.; Martinson, N.; Gray, G.; McIntyre, J.; Heneine, W. Emergence of drug-resistant HIV-1 after intrapartum administration of single-dose nevirapine is substantially underestimated. J. Infect. Dis. 2005, 192, 16–23. [Google Scholar] [CrossRef]
  136. Shafer, R.W. Rationale and uses of a public HIV drug-resistance database. J. Infect. Dis. 2006, 194, S51–S58. [Google Scholar] [CrossRef]
  137. McCormick, A.L.; Parry, C.M.; Crombe, A.; Goodall, R.L.; Gupta, R.K.; Kaleebu, P.; Kityo, C.; Chirara, M.; Towers, G.J.; Pillay, D. Impact of the N348I mutation in HIV-1 reverse transcriptase on nonnucleoside reverse transcriptase inhibitor resistance in non-subtype B HIV-1. Antimicrob. Agents Chemother. 2011, 55, 1806–1809. [Google Scholar] [CrossRef]
  138. Brehm, J.H.; Koontz, D.L.; Wallis, C.L.; Shutt, K.A.; Sanne, I.; Wood, R.; McIntyre, J.A.; Stevens, W.S.; Sluis-Cremer, N.; Mellors, J.W.; et al. Frequent emergence of N348I in HIV-1 subtype C reverse transcriptase with failure of initial therapy reduces susceptibility to reverse-transcriptase inhibitors. Clin. Infect. Dis. 2012, 55, 737–745. [Google Scholar] [CrossRef]
  139. Nikolenko, G.N.; Delviks-Frankenberry, K.A.; Palmer, S.; Maldarelli, F.; Fivash, M.J., Jr.; Coffin, J.M.; Pathak, V.K. Mutations in the connection domain of HIV-1 reverse transcriptase increase 3'-azido-3'-deoxythymidine resistance. Proc. Natl. Acad. Sci. USA 2007, 104, 317–322. [Google Scholar] [CrossRef]
  140. Vijayan, R.S.; Arnold, E.; Das, K. Molecular dynamics study of HIV-1 RT-DNA-nevirapine complexes explains NNRTI inhibition and resistance by connection mutations. Proteins 2014, 82, 815–829. [Google Scholar] [CrossRef]
  141. Delviks-Frankenberry, K.A.; Lengruber, R.B.; Santos, A.F.; Silveira, J.M.; Soares, M.A.; Kearney, M.F.; Maldarelli, F.; Pathak, V.K. Connection subdomain mutations in HIV-1 subtype-C treatment-experienced patients enhance NRTI and NNRTI drug resistance. Virology 2013, 435, 433–441. [Google Scholar] [CrossRef]
  142. Cane, P.; Chrystie, I.; Dunn, D.; Evans, B.; Geretti, A.M.; Green, H.; Phillips, A.; Pillay, D.; Porter, K.; Pozniak, A.; et al. Time trends in primary resistance to HIV drugs in the United Kingdom: Multicentre observational study. BMJ 2005, 331, 1368. [Google Scholar] [CrossRef]
  143. Shet, A.; Berry, L.; Mohri, H.; Mehandru, S.; Chung, C.; Kim, A.; Jean-Pierre, P.; Hogan, C.; Simon, V.; Boden, D.; et al. Tracking the prevalence of transmitted antiretroviral drug-resistant HIV-1: A decade of experience. J. Acquir. Immune Defic. Syndr. 2006, 41, 439–446. [Google Scholar] [CrossRef]
  144. Smith, D.; Moini, N.; Pesano, R.; Cachay, E.; Aiem, H.; Lie, Y.; Richman, D.; Little, S. Clinical utility of HIV standard genotyping among antiretroviral-naive individuals with unknown duration of infection. Clin. Infect. Dis. 2007, 44, 456–458. [Google Scholar] [CrossRef]
  145. Bennett, D.E.; Camacho, R.J.; Otelea, D.; Kuritzkes, D.R.; Fleury, H.; Kiuchi, M.; Heneine, W.; Kantor, R.; Jordan, M.R.; Schapiro, J.M.; et al. Drug resistance mutations for surveillance of transmitted HIV-1 drug-resistance: 2009 update. PLoS One 2009, 4, e4724. [Google Scholar] [CrossRef]
  146. Shafer, R.W.; Rhee, S.Y.; Bennett, D.E. Consensus drug resistance mutations for epidemiological surveillance: Basic principles and potential controversies. Antivir. Ther. 2008, 13, 59–68. [Google Scholar]
  147. Kantor, R.; Katzenstein, D. Polymorphism in HIV-1 non-subtype B protease and reverse transcriptase and its potential impact on drug susceptibility and drug resistance evolution. AIDS Rev. 2003, 5, 25–35. [Google Scholar]
  148. Martinez-Cajas, J.L.; Pant-Pai, N.; Klein, M.B.; Wainberg, M.A. Role of genetic diversity amongst HIV-1 non-B subtypes in drug resistance: A systematic review of virologic and biochemical evidence. AIDS Rev. 2008, 10, 212–223. [Google Scholar]
  149. Hirsch, M.S.; Brun-Vezinet, F.; Clotet, B.; Conway, B.; Kuritzkes, D.R.; D’Aquila, R.T.; Demeter, L.M.; Hammer, S.M.; Johnson, V.A.; Loveday, C.; et al. Antiretroviral drug resistance testing in adults infected with human immunodeficiency virus type 1: 2003 recommendations of an International AIDS Society-USA Panel. Clin. Infect. Dis. 2003, 37, 113–128. [Google Scholar] [CrossRef]
  150. Descamps, D.; Collin, G.; Letourneur, F.; Apetrei, C.; Damond, F.; Loussert-Ajaka, I.; Simon, F.; Saragosti, S.; Brun-Vezinet, F. Susceptibility of human immunodeficiency virus type 1 group O isolates to antiretroviral agents: In vitro phenotypic and genotypic analyses. J. Virol. 1997, 71, 8893–8898. [Google Scholar]
  151. Tuaillon, E.; Gueudin, M.; Lemee, V.; Gueit, I.; Roques, P.; Corrigan, G.E.; Plantier, J.C.; Simon, F.; Braun, J. Phenotypic susceptibility to nonnucleoside inhibitors of virion-associated reverse transcriptase from different HIV types and groups. J. Acquir. Immune Defic. Syndr. 2004, 37, 1543–1549. [Google Scholar] [CrossRef]
  152. Menendez-Arias, L.; Betancor, G.; Matamoros, T. HIV-1 reverse transcriptase connection subdomain mutations involved in resistance to approved non-nucleoside inhibitors. Antivir. Res. 2011, 92, 139–149. [Google Scholar] [CrossRef]
  153. Santos, A.F.; Silveira, J.; Muniz, C.P.; Tornatore, M.; Goes, L.R.; Mendoza-Sassi, R.A.; Martinez, A.M.; Tupinambas, U.; Greco, D.B.; Soares, M.A. Primary HIV-1 drug resistance in the C-terminal domains of viral reverse transcriptase among drug-naive patients from Southern Brazil. J. Clin. Virol. 2011, 52, 373–376. [Google Scholar] [CrossRef]
  154. Tanuma, J.; Hachiya, A.; Ishigaki, K.; Gatanaga, H.; Lien, T.T.; Hien, N.D.; Kinh, N.V.; Kaku, M.; Oka, S. Impact of CRF01_AE-specific polymorphic mutations G335D and A371V in the connection subdomain of human immunodeficiency virus type 1 (HIV-1) reverse transcriptase (RT) on susceptibility to nucleoside RT inhibitors. Microbes. Infect. 2010, 12, 1170–1177. [Google Scholar] [CrossRef]
  155. Muniz, C.P.; Soares, M.A.; Santos, A.F. Early selection of resistance-associated mutations in HIV-1 RT C-terminal domains across different subtypes: Role of the genetic barrier to resistance. J. Antimicrob. Chemother. 2014, 69, 2741–2745. [Google Scholar] [CrossRef]
  156. Hoffmann, C.J.; Charalambous, S.; Sim, J.; Ledwaba, J.; Schwikkard, G.; Chaisson, R.E.; Fielding, K.L.; Churchyard, G.J.; Morris, L.; Grant, A.D. Viremia, resuppression, and time to resistance in human immunodeficiency virus (HIV) subtype C during first-line antiretroviral therapy in South Africa. Clin. Infect. Dis. 2009, 49, 1928–1935. [Google Scholar] [CrossRef]
  157. El-Khatib, Z.; Ekstrom, A.M.; Ledwaba, J.; Mohapi, L.; Laher, F.; Karstaedt, A.; Charalambous, S.; Petzold, M.; Katzenstein, D.; Morris, L. Viremia and drug resistance among HIV-1 patients on antiretroviral treatment: A cross-sectional study in Soweto, South Africa. AIDS 2010, 24, 1679–1687. [Google Scholar] [CrossRef]
  158. Neogi, U.; Heylen, E.; Shet, A.; Chandy, S.; Shamsunder, R.; Sonnerborg, A.; Ekstrand, M.L. Long-term efficacy of first line antiretroviral therapy in Indian HIV-1 infected patients: A longitudinal cohort study. PLoS One 2013, 8, e55421. [Google Scholar] [CrossRef]
  159. Cihlar, T.; Ray, A.S. Nucleoside and nucleotide HIV reverse transcriptase inhibitors: 25 years after zidovudine. Antivir. Res. 2010, 85, 39–58. [Google Scholar] [CrossRef]
  160. Waters, L.J.; Moyle, G.; Bonora, S.; D'Avolio, A.; Else, L.; Mandalia, S.; Pozniak, A.; Nelson, M.; Gazzard, B.; Back, D.; et al. Abacavir plasma pharmacokinetics in the absence and presence of atazanavir/ritonavir or lopinavir/ritonavir and vice versa in HIV-infected patients. Antivir. Ther. 2007, 12, 825–830. [Google Scholar]
  161. Cahn, P.; Wainberg, M.A. Resistance profile of the new nucleoside reverse transcriptase inhibitor apricitabine. J. Antimicrob. Chemother. 2010, 65, 213–217. [Google Scholar] [CrossRef]
  162. Murphey-Corb, M.; Rajakumar, P.; Michael, H.; Nyaundi, J.; Didier, P.J.; Reeve, A.B.; Mitsuya, H.; Sarafianos, S.G.; Parniak, M.A. Response of simian immunodeficiency virus to the novel nucleoside reverse transcriptase inhibitor 4'-ethynyl-2-fluoro-2'-deoxyadenosine in vitro and in vivo. Antimicrob. Agents Chemother. 2012, 56, 4707–4712. [Google Scholar] [CrossRef]
  163. Michailidis, E.; Marchand, B.; Kodama, E.N.; Singh, K.; Matsuoka, M.; Kirby, K.A.; Ryan, E.M.; Sawani, A.M.; Nagy, E.; Ashida, N.; et al. Mechanism of inhibition of HIV-1 reverse transcriptase by 4'-Ethynyl-2-fluoro-2'-deoxyadenosine triphosphate, a translocation-defective reverse transcriptase inhibitor. J. Biol. Chem. 2009, 284, 35681–35691. [Google Scholar] [CrossRef]
  164. Muftuoglu, Y.; Sohl, C.D.; Mislak, A.C.; Mitsuya, H.; Sarafianos, S.G.; Anderson, K.S. Probing the molecular mechanism of action of the HIV-1 reverse transcriptase inhibitor 4'-ethynyl-2-fluoro-2'-deoxyadenosine (EFdA) using pre-steady-state kinetics. Antivir. Res. 2014, 106, 1–4. [Google Scholar] [CrossRef]
  165. Michailidis, E.; Huber, A.D.; Ryan, E.M.; Ong, Y.T.; Leslie, M.D.; Matzek, K.B.; Singh, K.; Marchand, B.; Hagedorn, A.N.; Kirby, K.A.; et al. 4'-Ethynyl-2-fluoro-2'-deoxyadenosine (EFdA) Inhibits HIV-1 Reverse Transcriptase with Multiple Mechanisms. J. Biol. Chem. 2014, 289, 24533–24548. [Google Scholar] [CrossRef]
  166. Kawamoto, A.; Kodama, E.; Sarafianos, S.G.; Sakagami, Y.; Kohgo, S.; Kitano, K.; Ashida, N.; Iwai, Y.; Hayakawa, H.; Nakata, H.; et al. 2'-deoxy-4'-C-ethynyl-2-halo-adenosines active against drug-resistant human immunodeficiency virus type 1 variants. Int. J. Biochem. Cell Biol. 2008, 40, 2410–2420. [Google Scholar] [CrossRef]
  167. Sohl, C.D.; Singh, K.; Kasiviswanathan, R.; Copeland, W.C.; Mitsuya, H.; Sarafianos, S.G.; Anderson, K.S. Mechanism of interaction of human mitochondrial DNA polymerase gamma with the novel nucleoside reverse transcriptase inhibitor 4'-ethynyl-2-fluoro-2'-deoxyadenosine indicates a low potential for host toxicity. Antimicrob. Agents Chemother. 2012, 56, 1630–1634. [Google Scholar] [CrossRef]
  168. Michailidis, E.; Ryan, E.M.; Hachiya, A.; Kirby, K.A.; Marchand, B.; Leslie, M.D.; Huber, A.D.; Ong, Y.T.; Jackson, J.C.; Singh, K.; et al. Hypersusceptibility mechanism of Tenofovir-resistant HIV to EFdA. Retrovirology 2013, 10, 65. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Singh, K.; Flores, J.A.; Kirby, K.A.; Neogi, U.; Sonnerborg, A.; Hachiya, A.; Das, K.; Arnold, E.; McArthur, C.; Parniak, M.; et al. Drug Resistance in Non-B Subtype HIV-1: Impact of HIV-1 Reverse Transcriptase Inhibitors. Viruses 2014, 6, 3535-3562. https://doi.org/10.3390/v6093535

AMA Style

Singh K, Flores JA, Kirby KA, Neogi U, Sonnerborg A, Hachiya A, Das K, Arnold E, McArthur C, Parniak M, et al. Drug Resistance in Non-B Subtype HIV-1: Impact of HIV-1 Reverse Transcriptase Inhibitors. Viruses. 2014; 6(9):3535-3562. https://doi.org/10.3390/v6093535

Chicago/Turabian Style

Singh, Kamalendra, Jacqueline A. Flores, Karen A. Kirby, Ujjwal Neogi, Anders Sonnerborg, Atsuko Hachiya, Kalyan Das, Eddy Arnold, Carole McArthur, Michael Parniak, and et al. 2014. "Drug Resistance in Non-B Subtype HIV-1: Impact of HIV-1 Reverse Transcriptase Inhibitors" Viruses 6, no. 9: 3535-3562. https://doi.org/10.3390/v6093535

Article Metrics

Back to TopTop