Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2009 Apr 1.
Published in final edited form as: AIDS Rev. 2008;10(2):67–84.

HIV-1 Drug Resistance Mutations: an Updated Framework for the Second Decade of HAART

Robert W Shafer 1, Jonathan M Schapiro 2
PMCID: PMC2547476  NIHMSID: NIHMS65073  PMID: 18615118

Abstract

More than 200 mutations are associated with antiretroviral resistance to drugs belonging to six licensed antiretroviral classes. More than 50 reverse transcriptase mutations are associated with nucleoside reverse transcriptase inhibitor resistance including M184V, thymidine analog mutations, mutations associated with non-thymidine analog containing regimens, multi-nucleoside resistance mutations, and several recently identified accessory mutations. More than 40 reverse transcriptase mutations are associated with nonnucleoside reverse transcriptase inhibitor resistance including major primary and secondary mutations, non-polymorphic minor mutations, and polymorphic accessory mutations. More than 60 mutations are associated with protease inhibitor resistance including major protease, accessory protease, and protease cleavage site mutations. More than 30 integrase mutations are associated with the licensed integrase inhibitor raltegravir and the investigational inhibitor elvitegravir. More than 15 gp41 mutations are associated with the fusion inhibitor enfuvirtide. CCR5 inhibitor resistance results from mutations that promote gp120 binding to an inhibitor-bound CCR5 receptor or CXCR4 tropism; however, the genotypic correlates of these processes are not yet well characterized.

Keywords: Reverse transcriptase, Protease, Integrase, Antiretrovirals, HIV-1 tropism, CCR5 inhibitors, HIV-1 mutations

Introduction

Nearly 25 antiretroviral drugs (ARV) have been licensed for the treatment of HIV-1: nine nucleoside reverse transcriptase inhibitors (NRTI), four nonnucleoside reverse transcriptase inhibitors (NNRTI), nine protease inhibitors (PI), one fusion inhibitor, one CCR5 inhibitor, and one integrase inhibitor. The first CCR5 and integrase inhibitors were approved in 2007, increasing the number of ARV classes from four to six. Commensurate with the increase in new ARV and ARV classes, there has been an increase in knowledge about drug resistance mutations. Entirely new vistas of mutations associated with integrase and CCR5 inhibitor resistance have also been opened, many new treatment-associated NRTI, NNRTI, and PI mutations have recently been described, and there has been a growing appreciation of the effects that different amino acid substitutions at the same position have on drug susceptibility.

Together with the expansion in the number of ARV classes and number of individual ARV, a consensus has emerged that ARV therapy can and should be used to completely suppress HIV-1 replication, even in patients in whom many previous ARV regimens have failed1,2. This unambiguous therapeutic endpoint (complete virologic suppression) necessitates a new framework in which the vast knowledge of drug resistance mutations should be cast. The identification of specific drug resistance mutations can increasingly be used to avoid ARV that retain only minimal residual activity in favor of newer ARV that are likely to be either fully or nearly fully active. Therefore, as the breadth of knowledge about HIV-1 drug resistance continues to expand, many of the subtle distinctions among drug resistance mutations are becoming less clinically relevant.

Nucleoside reverse transcriptase inhibitors

The NRTI resistance mutations include M184V, thymidine analog mutations (TAM), mutations selected by regimens lacking thymidine analogs, multi-nucleoside resistance mutations, and many recently described non-polymorphic accessory mutations. There are two biochemical mechanisms of NRTI resistance: enhanced discrimination against and decreased incorporation of NRTI in favor of authentic nucleosides, and enhanced removal of incorporated NRTI by promoting a phosphorolytic reaction that leads to primer unblocking. Altogether, M184V, non thymidine analog-associated mutations such as K65R and L74V, and the multi-nucleoside resistance mutation Q151M act by decreasing NRTI incorporation3,4. Thymidine analog mutations, the T69 insertions associated with multi-nucleoside resistance, and many of the accessory mutations facilitate primer unblocking5,7.

M184V

M184V is the most commonly occurring NRTI resistance mutation. In vitro, it causes high-level resistance to lamivudine (3TC) and emtricitabine (FTC), low-level resistance to didanosine (ddI) and abacavir, (ABC) and increased susceptibility to zidovudine (ZDV), stavudine (d4T), and tenofovir (TDF)8. The possibility that isolates with M184V are compromised was suggested by the initial 3TC monotherapy studies showing that plasma HIV-1 RNA levels remained about 0.5 log10 copies below baseline in patients receiving lamivudine for 6-12 months, despite the development of M184V and a high level of phenotypic resistance to 3TC9-11. Data from multiple 3TC-containing dual-NRTI regimens also suggest that 3TC continues to exert an antiviral effect even in patients whose virus isolates contain M184V12-14.

M184V causes a median 1.5-fold and 3.0-fold reduction in susceptibility to ddI and ABC, respectively, in the PhenoSenseGT™ assay (Monogram Biosciences)15,16. These are levels of reduction that are above the wild-type range but below the level at which these NRTI are inactive15. Several clinical trials have also shown that ABC and ddI retain clinical activity in the presence of M184V17-22. For example, the addition of ddI or ABC to the regimen of a patient with virologic failure has been associated to plasma HIV-1 RNA reductions of 0.6 and 0.7 log10, respectively, in patients harboring viruses with M184V and no other drug resistance mutations19,21. The phenotypic and clinical significance of M184V is influenced by the presence or absence of other NRTI resistance mutations. For example, the presence of K65R or L74V in combination with M184V is sufficient for high-level resistance to both ABC and ddI16. In contrast, three or more TAM plus M184V are required for high-level ABC and ddI resistance8,16,19,20,23.

Thymidine analog mutations

Thymidine analog mutations are selected by the thymidine analogs ZDV and d4T. Thymidine analog mutations decrease susceptibility to these NRTI and to a lesser extent to ABC, ddI, and TDF8. Thymidine analog mutations are common in low-income countries in which fixed-dose combinations containing thymidine analogs are the mainstays of therapy. Thymidine analog mutations are also common in viruses from persons who began therapy in the pre-HAART era with incompletely suppressive thymidine analog-containing regimens, but are becoming less common in areas in which the fixed-dose combinations of TDF/FTC and ABC/3TC have become the most common NRTI backbones. However, even in these areas, TAM and in particular the partial T215 revertants remain the most common type of transmitted NRTI resistance mutation24,25 (Table 1).

Table 1. Nucleoside reverse transcriptase inhibitor resistance mutations*.

NRTI 184 Thymidine analog mutations (TAM) Non thymidine analog regimen mutations Multi-NRTI resistance mutations

41 67 70 210 215 219 65 70 74 75 115 69 151 62 75 77 116

M M D K L T K K K L V Y T Q A V F F
3TC VI RN EG Ins M V
FTC VI RN EG Ins M V
ABC VI L N W FY RN EG VI TM F Ins M V I L Y
DDI VI L N W FY RN EG VI TM Ins M V I L Y
TDF L N W FY RN EG M F Ins M V
D4T L N R W FY QE RN TM Ins M V I L Y
ZDV L N R W FY QE Ins M V I L Y
*

The first row of letters contains the consensus amino acid at the position indicated by the number in the preceding row. All amino acids are indicated by their one letter code with the exception of “Ins” which is an abbreviation for one or more amino acid insertions. Mutations in bold are associated with higher levels of phenotypic resistance or clinical evidence for reduced virologic response. Additional treatment-selected mutations at the positions in this table include D67G/E, T69DS/A/I/N/G, K70N, V75A/S, and K219NR. Additional accessory mutations include K43E/Q/N, E44D/A, V118I, H208Y, D218E, H221Y, and L228H/R. These accessory mutations generally occur with TAM and appear to be associated with a reduced level of susceptibility to multiple NRTI. Several mutations are associated with increased susceptibility: M184V/I increases susceptibility to ZDV, TDF, and d4T; L74V increases susceptibility to ZDV and TDF; K65R increases susceptibility to ZDV.

K70R occurs in viruses from patients receiving thymidine analogs; K70E/G occur with non thymidine analog-containing regimens.

V75I occurs in combination with Q151M; V75TM occur in a variety of different treatment and mutational contexts.

NRTI: nucleoside reverse transcriptase inhibitor; 3TC: lamivudine; FTC: emtricitabine; ABC: abacavir; DDI: didanosine; TDF: tenofovir; D4T: stavudine; ZDV: zidovudine.

Thymidine analog mutations accumulate in two distinct but overlapping patterns26-31. The type I pattern includes the mutations M41L, L210W, and T215Y. The type II pattern includes D67N, K70R, T215F, and K219Q/E. Mutation D67N also occurs commonly with type I TAM30,32. However, K70R and L210W rarely occur together33. Type I TAM cause higher levels of phenotypic and clinical resistance to the thymidine analogs and cross-resistance to ABC, ddI, and TDF than do the type II TAM. Indeed, the presence of all three type I TAM markedly reduces the clinical response to ABC, ddI, and TDF19,29,30,34,35. The clinical significance of the type II TAM is not as well characterized.

Other mutations at several of the TAM positions are common. The most common of these are the partial T215 revertants T215C/D/E/I/S/V36,37. These mutations arise from the drug resistance mutations T215Y/F to increase HIV-1 fitness in the absence of selective drug pressure. They occur more commonly than reversion to the wild-type T because most of the partial T215 revertants require only a single nucleotide mutation rather than the double nucleotide mutation required for Y or F to revert to T. The partial T215 revertants do not reduce drug susceptibility by themselves, but their presence in a previously untreated patient suggests that the patient may have been infected originally with a virus containing T215Y or F. Both K219N/R are two variants that unlike K219Q/E usually occur with type I rather than type II TAM32. Interestingly, two variants at position 70, K70E/G, are not selected by thymidine analogs and have phenotypic effects diametrically opposite to those of K70R, decreasing ABC, ddI, TDF, 3TC, and FTC susceptibility, and increasing ZDV susceptibility38-40. Both D67G and D67E are selected by NRTI therapy, but their phenotypic and clinical significance are not well characterized41.

E44D/A and V118I are accessory mutations that generally occur with type I TAM. These mutations occur in about 1% of viruses from untreated patients and in a significantly higher proportion of viruses from patients receiving NRTI27,42,43. Although E44D plus V118I were first shown to cause low-level 3TC resistance when they occur in combination44, subsequent studies have suggested that in combination with TAM, these mutations reduce the susceptibility and clinical activity of most NRTI20,43,45-52. F214L is a common polymorphism that is negatively associated with type I TAM, and as a consequence may raise the genetic barrier to resistance in viruses developing type I TAM53,54.

Mutations occurring in the absence of thymidine analogs

The most common mutations in patients developing virologic failure while receiving a non thymidine analog-containing NRTI backbone include M184V alone or M184V in combination with K65R or L74V55-57. K65R causes intermediate resistance to TDF, ABC, ddI, 3TC, and FTC, low-level resistance to d4T, and increased susceptibility to ZDV58-60. L74V causes intermediate resistance to ddI and ABC, and a slight increase in susceptibility to ZDV and TDF61. L74I has similar phenotypic properties to L74V, but is found primarily in viruses with multiple TAM, possibly because it increases ZDV and TDF susceptibility less than L74V62,63.

Mutations M184V plus K65R have been reported primarily in patients receiving the NRTI backbone TDF/3TC64-65 and less commonly ABC/3TC55,66 or TDF/ FTC56,67. M184V plus L74V occurs primarily in persons receiving ABC/3TC or ddI/3TC/FTC backbones55,66,68,69. K65R and L74V rarely occur in the same viruses; however, several patients developing virologic failure with L74V while receiving an ABC- or ddI-containing regimen have been found to have minor variants containing K65R69,70.

There is a bidirectional antagonism between K65R and the TAM. K65R interferes with TAM-mediated primer unblocking and the TAM interfere with K65R-mediated NRTI discrimination71,72. As a result, viruses containing K65R in combination with TAM are uncommon73. The emergence of K65R is suppressed to a greater extent in regimens containing ZDV compared with d4T59,74-81.

Less common mutations occurring during virologic failure with non thymidine analog regimens include K65N, K70E/G, and Y115F38,40,55,82,83. K65N and K70E/G have a resistance profile similar to K65R, but appear to cause less resistance than K65R to ABC, ddI, TDF, 3TC, and FTC38-40,82,84,85. Y115F reduces ABC susceptibility86 and causes low-level cross-resistance to TDF23,58,61,87. Although T69D and V75T were originally identified as causing resistance to ddC88 and d4T89, respectively, a range of mutations at these positions (e.g. T69N/S/I/G and V75M/A) have been associated with reduced susceptibility to other NRTI, including ddI and d4T23,89-92.

Two lines of evidence suggest that K65R may occur more commonly in non subtype B compared with subtype B viruses. K65R has emerged more rapidly during the in vitro passage of subtype C compared with subtype B isolates in the presence of increasing TDF concentrations93. Anecdotal reports have also suggested that K65R may occur more commonly in low-income countries when patients with non-B subtype viruses are treated with d4T/ddI and d4T/3TC94,95, or TDF/3TC96.

Multi-nucleoside resistance mutations

Amino acid insertions at codon 69 generally occur in the presence of multiple TAM, and in this setting are associated with intermediate resistance to 3TC and FTC and high-level resistance to each of the remaining NRTI97-101. Q151M is a 2-bp mutation (CAG→ATG) that is usually accompanied by two or more of the following mutations: A62V, V75I, F77L, and F116Y. The Q151M complex causes high-level resistance to ZDV, d4T, ddI, and ABC, and intermediate resistance to TDF, 3TC, and FTC61,102,103. This complex developed in 5% of patients who received ddI in combination with ZDV or d4T98,104, but is rarely selected by 3TC- or FTC-containing regimens. Q151M may be uncommon because the two intermediate amino acids Q151L (CAG→CTG) and Q151K (CAG→AAG) are poorly replicating and rarely observed105-107. Q151M is a common genetic mechanism of NRTI resistance in HIV-2-infected persons108,109. The optimal NRTI combination to use in patients with codon 69 insertions or Q151M is not known110,111.

Miscellaneous mutations

Mutations K43E/Q/N, E203D/K, H208Y, D218E, H221Y, K223Q, and L228H/R are non-polymorphic NRTI-selected mutations which generally follow TAM and which have subtle effects on HIV-1 NRTI susceptibility and replication27,53,61,112. Q145M is a rare mutation that has been reported by one group to reduce susceptibility to multiple NRTI and NNRTI113,114. P157S, which is homologous to the mutation causing 3TC resistance in FIV, has been reported once in an HIV-1 isolate115,116.

Several mutations in the connection and RNaseH domains of HIV-1 RT play an accessory role in reducing HIV-1 susceptibility in combination with TAM, most likely by slowing the activity of RNaseH and thereby allowing more time for TAM-mediated primer unblocking117. The single most important of these mutations may be N348I, a non-polymorphic mutation that occurs in about 10% of NRTI-treated patients118. N348I causes a twofold reduction in ZDV susceptibility when it occurs in combination with multiple TAM118. G333E/D, A360T, and A371V, mutations with similar phenotypic effects, occur in about 5% of NRTI-naive and 10% of NRTI-treated patients119-122. Although several RNaseH mutations may potentially reduce ZDV susceptibility in combination with TAM123, few have been observed in clinical isolates124,125.

Nonnucleoside reverse transcriptase inhibitors

The NNRTI inhibit HIV-1 RT allosterically by binding to a hydrophobic pocket close to but not contiguous with the RT active site. Nearly all of the NNRTI resistance mutations are within the NNRTI binding pocket or adjacent to residues in the pocket126,127. There is a low genetic barrier to NNRTI resistance, with only one or two mutations required for high-level resistance. High levels of clinical cross-resistance exist among the NNRTI because many of the NNRTI resistance mutations reduce susceptibility to multiple NNRTI and because the low genetic barrier to resistance allows a single NNRTI to select for multiple NNRTI resistance mutations in different viruses, even if only a single mutation is detected by standard population-based sequencing128,129.

The NNRTI resistance mutations can be classified into the following categories: (i) primary NNRTI resistance mutations that cause high-level resistance to one or more NNRTI and that are among the first to develop during NNRTI therapy; (ii) secondary NNRTI resistance mutations that usually occur in combination with primary NNRTI resistance mutations, but that also have clinically significant implications for choosing an NNRTI, particularly etravirine; (iii) minor non-polymorphic mutations that may occur alone or in combination with other NNRTI resistance mutations and that cause consistent but low-level reductions in NNRTI susceptibility; and (iv) polymorphic accessory mutations that modulate the effects of other NNRTI resistance mutations. Table 2 summarizes effect of the major primary, major secondary, and minor NNRTI resistance mutations on delavirdine, efavirenz, etravirine, and nevirapine.

Table 2. Nonnucleoside reverse transcriptase inhibitor resistance mutations*.

98 100 101 103 106 108 179 181 188 190 225 227 230 236 238

A L K K V V V Y Y G P F M P K
NVP G I EP NS AM I DEF CIV LHC ASE LC L NT
DLV G I EP NS AM I DEF CIV LHC E C L L NT
EFV G I EP NS AM I DEF CIV LHC ASE H C L NT
ETR G I EP DEF CIV LHC ASE C L
*

The first row of letters contains the consensus amino acid at the position indicated by the number in the preceding row. All amino acids are indicated by their one letter code. Mutations in bold are associated with higher levels of phenotypic resistance or clinical evidence for reduced virologic response. Several additional uncommon mutations at the positions in this table are also associated with NNRTI therapy or phenotypic resistance including: K101N/H, K103T/H, G190Q/C/T/V. Additional NNRTI resistance mutations that are not in the table include E138K, L234I, and L318F. E138K has been selected in vitro by ETR and been shown to cause low-level reductions in susceptibility to each of the NNRTI133,145. L234I has been selected in vitro by ETR, acts synergistically with Y181C to reduce ETR susceptibility133. L318F is a non-polymorphic NNRTI-selected mutation that decreases susceptibility to DLV and to a lesser extent to nevirapine and possibly ETR133,146. Several polymorphic mutations such as K101Q, I135T/M, V179I, and L283I and NRTI selected mutations such as L74V, H221Y and N348I may cause subtle reductions in NNRTI susceptibility118,152. A98S, K101R/Q, K103R, V106I, E138A, V179I, and K238R are polymorphic substitutions with little if any effect on drug resistance on their own. However, the combination of K103R + V179D (each of which occurs in 1-2% of untreated persons) reduces susceptibility to NVP, DLV, and EFV about 15-fold141.

In the DUET study, a univariate analysis showed that persons with viruses with three or more of the following mutations responded similarly to placebo and ETR: V90I, A98G, L100I, K101E/P, V106I, V179D/F, Y181C/I/V, and G190A/S139.

NNRTI: nonnucleoside reverse transcriptase inhibitor; NVP: nevirapine; DLV: delavirdine; EFV: efavirenz; ETR: etravirine.

Because delavirdine is rarely used, it is not discussed in the sections that follow. The resistance profile of delavirdine is distinguished from that of the other NNRTI by the fact that G190A/S increase delavirdine susceptibility, providing perhaps the only virologic rationale for its use130. Although there have been case reports of virologic responses to delavirdine-containing salvage treatment regimens in treating viruses with G190A/S, the extent to which delavirdine contributed to these successes is not known131.

Primary nonnucleoside reverse transcriptase inhibitor resistance mutations

Each of the primary NNRTI resistance mutations – K103N/S, V106A/M, Y181C/I/V, Y188L/C/H, and G190A/S/E – cause high-level resistance to nevirapine and variable resistance to efavirenz, ranging from about twofold for V106A and Y181C, sixfold for G190A, 20-fold for K103N, and more than 50-fold for Y188L and G190S61,132,133. Although transient virologic responses to an efavirenz-based salvage therapy regimen occur in some NNRTI-experienced patients, a sustained response has been uncommon128,134-136. In contrast, patients with any single one of the primary NNRTI resistance mutations may benefit from etravirine salvage therapy, although the mutations at position 181 and to a lesser extent 190 compromise etravirine response and may provide the foundation for the development of high-level etravirine resistance137-139.

Major secondary nonnucleoside reverse transcriptase inhibitor resistance mutations

L100I, K101P, P225H, F227L, M230L, and K238T are secondary mutations that usually occur in combination with one of the primary NNRTI resistance mutations. L100I and K101P, which occur in combination with K103N, further decrease nevirapine and efavirenz susceptibility from 20-fold with K103N alone to more than 100-fold61. Although viruses with K103N are fully susceptible to etravirine, viruses with L100I plus K103N display about 10-fold decreased susceptibility133. P225H and K238T/N usually occur in combination with K103N and synergistically reduce nevirapine and efavirenz susceptibility132,140,141. F227L nearly always occurs in combination with V106A, leading to synergistic reductions in nevirapine susceptibility142. M230L, which may occur alone, decreases the susceptibility of all NNRTI including etravirine by 20-fold or more133,143.

V179F, F227C, L234I, and L318F are rare mutations that are of increased importance now that etravirine is licensed. V179F occurs solely in combination with Y181C/I/V and acts synergistically to increase etravirine resistance from fivefold to 10-fold with Y181C/I/V alone to more than 100-fold133. F227C, an exceedingly rare mutation, reduces etravirine susceptibility 10-fold to 20-fold144,145. L234I, which has been selected in vitro by etravirine, acts synergistically with Y181C to reduce etravirine susceptibility133. L318F, which was first reported to reduce delavirdine and nevirapine susceptibility by 15-fold and threefold, respectively146, has also been selected in vitro by etravirine and found to reduce etravirine susceptibility synergistically with Y181C133.

Minor nonnucleoside reverse transcriptase inhibitor resistance mutations

A98G, K101E, V108I, and V179D/E are common NNRTI resistance mutations that reduce susceptibility to nevirapine and efavirenz about twofold to fivefold147. Although K103R alone, which occurs in about 1% of untreated persons, has no effect on NNRTI susceptibility, the combination of K103R plus V179D reduces nevirapine and efavirenz susceptibility by 15-fold141. Data are not available on the effect of these mutations on etravirine susceptibility. V179D, and rarely A98G and V108I, are observed in patients who have never been treated with NNRTI148. The optimal management of patients with viruses containing these mutations is not known. Although low-level baseline resistance has not been shown to decrease the virologic responses to first-line NNRTI-containing regimens149, efavirenz and etravirine may be preferable to nevirapine because these NNRTI have generally been more active than nevirapine against these and other NNRTI-resistant variants127,150.

Miscellaneous nonnucleoside reverse transcriptase inhibitor resistance mutations

Several highly polymorphic RT mutations, such as K101Q, I135T/M, V179I, and L283I, reduce susceptibility to nevirapine and efavirenz by about twofold and may act synergistically with primary NNRTI resistance mutations151,152. Other mutations such as L74V, H221Y, K223E/Q, L228H/R, and N348I are selected primarily by NRTI, yet also cause subtle reductions in NNRTI susceptibility41,107,112,118,152-154. V90I and V106I are highly polymorphic mutations that were associated with decreased virologic response to etravirine in the DUET clinical trial, but may owe this association to their correlation with other NNRTI resistance mutations139. Mutations at positions 31, 135, and 245 have been reported to cause low-level NNRTI resistance in a non subtype B context155,156. Conversely, there is a large body of evidence showing that type I TAM increase NNRTI susceptibility157,158.

Protease inhibitors

As the PI class has expanded to nine licensed ARV, the individual PI have evolved increasingly specific roles. Ritonavir is used solely for pharmacokinetic boosting (indicated by/r). Lopinavir/r, atazanavir/r, fosamprenavir/r, and less commonly saquinavir/r are used for first-line therapy, whereas lopinavir/r, tipranavir/r, and darunavir/r are used for salvage therapy1,2. Nelfinavir, which cannot be boosted by ritonavir, and unboosted atazanavir and fosamprenavir are alternative but suboptimal choices for first-line therapy because of their higher risk of virologic failure with drug resistance compared with boosted PI. Although indinavir/r may be effective for first-line or salvage therapy, it is not recommended because of its high risk of nephrolithiasis.

More mutations are selected by the PI than by any other ARV class. The effect of PI resistance mutations on individual PI may be difficult to quantify when many mutations are present in the same virus isolate or when mutations occur in unusual patterns. The effect of PI resistance mutations on drug susceptibility can also be modulated by gag cleavage site mutations and possibly other parts of gag that influence Gag-Pol processing. Although multiple protease mutations are often required for HIV-1 to develop clinically significant resistance to a ritonavir-boosted PI159-161, some mutations indicate that a particular PI, even when boosted, may not be effective. Many protease mutations are accessory, compensating for the replication impairment of other PI resistance mutations or reducing PI susceptibility only in combination with other PI resistance mutations.

Major protease inhibitor resistance mutations

Table 3 lists mutations at 17 largely non-polymorphic positions that are of the most clinical significance. Mutations at 13 of these 17 positions have been shown to reduce susceptibility to one or more PI, including mutations at the substrate cleft positions 23, 30, 32, 47, 48, 50, 82, and 84, the flap positions 46 and 54, and interior enzyme positions 76, 88, and 90. Mutations at four of these 17 positions (24, 33, 53, and 73) are included because they are non-polymorphic, occur commonly, and have disparate effects on different PI61.

Table 3. Protease inhibitor resistance mutations*.

23 24 30 32 33 46 47 48 50 53 54 73 76 82 84 88 90

L L D V L M I G I F I G L V I N L
ATVr I F IL V VM L L VTALM ST ATFS VAC DS M
DRVr I F VA V LM ST V VAC M
FPVr I F IL VA V VTALM ST V ATFS VAC M
IDVr I V IL V L VTALM ST V AFTS VAC S M
LPVr I I F IL VA VM V VTALM V AFTS VAC M
NFV I I N F IL V VM L VTALM ST AFTS VAC DS M
SQVr I VM L VTALM ST AT VAC S M
TPVr§ I F IL V VAM ATFSL VAC M
*

The first row of letters contains the consensus amino acid at the position indicated by the number in the preceding row. All amino acids are indicated by their one letter code. Mutations in bold have been shown to reduce in vitro susceptibility or in vivo virologic response. Mutations in bold underline are relative contraindications to the use of specific PI. Several additional uncommon mutations at the 17 positions in this table are also selected by PI, but have not been evaluated phenotypically including L24F, L33I, M46V, F53Y, I54S, G73C/A, V82M/C, and N88T/G. In contrast, V82I and L33V are polymorphisms that are not associated with PI therapy. Accessory protease mutations that are not in the table include the polymorphic mutations L10I/V, I13V, K20R/M/I, M36I, D60E, I62V, L63P, A71V/T, V77I, and I93L and the non-polymorphic mutations L10F/R, V11I, E34Q, E35G, K43T, K45I, K55R, Q58E, A71I/L, T74P/A/S, V75I, N83D, P79A/S, I85V, L89V, T91S, Q92K and C95F.

I50L increases susceptibility to all PI except ATV; I50V and I54L increase TPV susceptibility; N88S increases FPV susceptibility; L76V increases ATV, SQV and TPV susceptibility.

A genotypic susceptibility score (GSS) for DRV based on the POWER clinical trials includes the number of the following 11 mutations: V11I, V32I, L33F, I47V, I50V, I54L/M, G73S, L76V, I84V, and L89V201. In a subsequent update the substitution of T74P for G73S led to an improved model202.

§

A GSS for TPV/r based on the RESIST studies identified 21 mutations at 11 positions: L10V, I13V, K20M/R/V, L33F, E35G, M36I, K43T, M46L, I47V, I54A/M/V, Q58E, H69K, T74P, V82L/T, N83D, and I84V184. An updated TPV/r GSS excluded I13V, K20M/R/V, E35G, and H69K; reclassified I47V, I54A/M/V, Q58E, T74P, V82L/T, and N83D as major mutations; reclassified L10V, M36I, K43T, M46L, and I84V as minor mutations; and included L24I, I50L/V, I54L, and L76V as mutations likely to improve TPV susceptibility and virological response200. A complete listing of studies of genotypic PI response predictors can be found at: http://hivdb.stanford.edu/pages/geno_clinical_review/PI.html ATVr: atazanavir/ritonavir(r); DRVr: darunavir/r; FPVr: fosamprenavir/r; IDVr: indinavir/r; LPVr: lopinavir/r; NFV: nelfinavir; SQVr: saquinavir/r; TPVr: tipranavir/r.

Whereas many mutations reduce nelfinavir susceptibility, L23I, D30N, M46I/L, G48V/M, I84V, N88D/S, and L90M are relative contraindications to the use of nelfinavir in that an inferior virologic response to therapy relative to that obtainable with most other PI would be expected14,162-167. I50L and N88S and possibly I84V, are relative contraindications for the use of atazanavir/r23,61,168-174. G48V/M, I84V, and L90M are relative contraindications to the use of saquinavir/r175-177. V32I, I47V/A, I54L/M, and I84V are relative contraindications to the use of fosamprenavir/r174,178-181. Mutations at position 82 as well as I84V may be relative contraindications to the use of indinavir/r. There are few known contraindications to the salvage therapy PI (lopinavir/r, tipranavir/r, darunavir/r), except V47A for lopinavir/r178,182,183 and V82L/T for tipranavir/r184.

At six of the 17 PI resistance mutations in table 3, only a single mutation has been shown to be associated with PI resistance – L23I, L24I, D30N, V32I, L76V, and L90M. At 11 positions, different mutations are associated with PI resistance, and at positions 50, 54, 82, and 88 these differences can be responsible for dramatically different effects on PI susceptibility. Additional, uncommon, PI-selected mutations not shown in table 3 include L33I, M46V, F53Y, I54S, G73C/A, V82M/C, and N88T/G23,41,185. V82I, which does not contribute to PI resistance, is a polymorphism that is the consensus residue for subtype G isolates. L33V is another polymorphism that is not associated with PI therapy or resistance. L33F and M46I/L, although non-polymorphic in most subtypes, occur at a prevalence of about 0.5-1% in subtype A and CRF01_AE isolates (http://hivdb.stanford.edu/cgi-bin/MutPrevBySubtypeRx.cgi)148.

Several resistance mutations are associated with increased susceptibility to one or more PI, including I50L which increases susceptibility to all PI other than atazanavir168, I50V and I54L which increase tipranavir susceptibility186, N88S which increases fosamprenavir susceptibility187, and L76V which increases susceptibility to atazanavir, saquinavir, and tipranavir23,188.

Accessory protease inhibitor resistance mutations

Mutations at positions 10, 20, 36, 63, and 71 up-regulate protease processivity to compensate for the decreased fitness associated with the major PI resistance mutations189-193. Positions 20, 36, and 63 are highly polymorphic. In contrast, L10I/V and A71V/T occur in 5 and 10%, respectively, of PI-naive patients, and in a much higher proportion of PI-treated patients, while L10F/R and A71I/L do not occur in the absence of PI therapy147. In one retrospective study, baseline mutations at positions 10 and 36 were associated with an increased risk of virologic failure in patients receiving older PI-based regimens containing nelfinavir or an unboosted PI194,195.

Additional PI-selected accessory mutations include the highly polymorphic mutations I13V, D60E, I62V, V77I and I93L, and many uncommon non-polymorphic mutations including V11I, E34Q, E35G, K43T, K45I, K55R, Q58E, T74P/A/S, V75I, N83D, P79A/S, I85V, L89V, T91S, Q92K and C95F23,41,196-199. Several of the non-polymorphic mutations have become part of the genotypic susceptibility scores for tipranavir/r (E35G, K43T, Q58E, T74P, and N83D) and darunavir (V11I, T74P, and L89V), based on analyses of the RESIST184,200 and POWER and DUET201,202 clinical trials. These mutations, however, have not been evaluated for their effects on other PI, but their presence at baseline in these two clinical trials for heavily PI-experienced patients suggests that they are also associated with decreased susceptibility to the older PI.

Gag cleavage site mutations

The gag gene codes for the matrix (MA), capsid (CA), and nucleocapsid (NC) proteins, a protein of uncertain function, p6, and two spacer peptides: p2 (between CA and NC) and p1 (between NC and p6). The gag polypeptide is cleaved at the MA/CA, CA/p2, p2/NC, NC/p1, and p1/6 junctions. A stem-loop structure between p1 and p6 stimulates the frame shifting necessary to create the Gag-Pol polypeptide. The residues surrounding each protease cleavage site are designated 5′-P4, P3, P2, P1/P1′, P2′, P3′, P4′-3′.

Mutations that improve the kinetics of PI-resistant proteases emerge at several protease cleavage sites during PI treatment203-205. Most gag cleavage site mutations occur at NC/p1 and p1/p6203,206 – sites at which cleavage may be rate limiting for gag and Gag-Pol polyprotein processing207. A431V, at the P2 position of NC/p1, is associated with mutations at protease positions 24, 46, and 82208,209. L449F, at the P1′ position of p1/p6, is associated with the protease mutation pair D30N/ N88D and with I84V209,210. P453L, at the P5′ position of the p1/p6 site, is associated with protease mutations at positions 32211, 47211, 50212, 84, and 90209,213. A set of three NC/p1 mutations (A431V, K436E, and I437T/V) developing during in vitro selection with the investigational PI RO033-4649 was found to cause a twofold reduction in susceptibility to multiple PI, even in the absence of mutations in protease214. Several p6 mutations, including insertions in a proline-rich region containing a conserved PTAP motif, occur more frequently in viruses with PI resistance mutations than in wild-type viruses215-219.

Subtype-specific mechanisms of protease inhibitor resistance

Naturally occurring polymorphisms in the different protease subtypes often occur at sites of accessory PI resistance mutations in subtype B isolates220. For example, the accessory PI resistance mutations I13V, K20I, M36I, and I93L represent the consensus variant in one or more non-B subtypes221. Although these mutations may result in subtle structural and biochemical differences among subtypes222-224, the vast majority of in vitro and in vivo studies suggest that the licensed PI are as active against wild-type non-B viruses as they are against wild-type subtype B viruses220,225.

With several notable exceptions, the genetic mechanisms of PI resistance are also highly similar among the different subtypes226. Although both D30N and L90M occur in non-B viruses during nelfinavir therapy, D30N occurs more commonly in subtype B viruses and L90M occurs more commonly in subtype C, F, G, and CRF01_AE viruses227-231. The increased predilection for certain subtypes to develop L90M may relate to the presence of variants other than L (the subtype B consensus) at position 89230-232. Similarly, T74S, a polymorphism that occurs in 8% of subtype C sequences, but rarely in other subtypes, is associated with reduced susceptibility to nelfinavir61,233.

The fact that V82I is the consensus amino acid for subtype G affects the spectrum of mutations observed at this position in PI-resistant subtype G isolates: V82T and the rare mutation V82M occur more frequently than V82A in subtype G isolates because T and M require a single base pair change, whereas A requires two base pair changes234. However, for nearly all other subtypes and protease mutations, a similar number of nucleotide changes is required to convert a wild-type residue into one associated with drug resistance235.

Integrase inhibitors

The HIV-1 integrase contains 288 amino acids encoded by the 3′ end of the HIV-1 pol gene. It has three functional domains: the N-terminal domain (NTD), which encompasses amino acids 1-50 and contains an HHCC motif that coordinates zinc binding236, the catalytic core domain (CCD) which encompasses amino acids 51-212 and contains the catalytic triad D64, D116, and E152, known as the DDE motif, and the C-terminal domain (CTD), which encompasses amino acids 213-288 and is involved in host DNA binding through an as yet poorly defined mechanism.

A multimeric form of integrase catalyzes the cleavage of the conserved 3′ dinucleotide CA (3′ processing) and the ligation of the viral 3′-OH ends to the 5′-DNA of host chromosomal DNA (strand transfer)237. Crystal structures of the CCD plus CTD domains238 and the CCD plus NTD239 have been solved, but the relative conformation of the three sub domains and of the active multimeric form of the enzyme are not known. There has been one published crystal structure of the CCD bound to an early prototype inhibitor (5CITEP)240 but no structures of the CCD bound to one of the integrase inhibitors (INI) in clinical use or to a DNA template.

The current generation of clinically relevant INI (the FDA-licensed inhibitor raltegravir and the investigational inhibitor elvitegravir) preferentially inhibit strand transfer by binding to the target DNA site of the enzyme. These INI as well as the initial series of strand-transfer diketo acid inhibitors including S-1360241 and L870,810242 select for mutations in the part of the integrase bound to 5CITEP240,243,244. In vitro drug susceptibility data and surveys of integrase sequences from HIV-1-infected patients previously treated with other ARV classes or who were treatment-naive suggest that there is no cross-resistance between the INI and the other HIV-1 enzyme inhibitors245-248.

Most INI resistance mutations are in the vicinity of the putative INI binding pocket. Some of the INI resistance mutations decrease susceptibility by themselves, whereas others compensate for the decreased fitness associated with other INI resistance mutations249. There is a high level of cross-resistance between raltegravir and elvitegravir, as well as between these INI and the first generation of strand-transfer inhibitors, suggesting that the development of non cross-resistant INI will be challenging245,250-255.

Among 38 patients with virologic failure in Merck Protocol 005, nearly all developed INI resistance mutations including N155H or Q148H/R/K, each of which reduces raltegravir susceptibility by 10-fold to 25-fold251. Higher levels of raltegravir resistance occurred with the accumulation of additional mutations. E92Q and the two polymorphic mutations L74M and G163R generally occurred with N155H, whereas G140A/S generally occurred with Q148H/R/K251. Additional mutations reported to the FDA as being selected either in vitro or in vivo by raltegravir include the non-polymorphic mutations L74R, E138A/K, Y143R/C/H, N155S, H183P, Y226D/F/H, S230R, and D232N and the polymorphic mutations T97A and V151I256,257.

Among 30 patients developing virologic failure while receiving elvitegravir in GS-US-1830105, 28 developed INI resistance mutations including E92Q, E138K, Q148H/R/K, or N155H in 11 patients, and S147G or T66I/A/K in nine and five patients, respectively252. Additional mutations selected in vitro by elvitegravir include the non-polymorphic mutations H51Y, Q95K, F121Y, Q146P, S153Y, and R263K, and the slightly polymorphic mutation E157Q245,250. For both raltegravir and elvitegravir, virologic failure has generally been accompanied by 100-fold or greater decreases in susceptibility and the development of two or more INI resistance mutations.

Table 4 lists the non-polymorphic INI resistance mutations that have been selected in patients receiving raltegravir or elvitegravir, or that have been characterized in vitro for susceptibility to both drugs. Mutations at positions 92, 121, 140, 148, and 155 are associated with more than fivefold to 10-fold decreased susceptibility to both INI, whereas mutations at positions 66 and 147 are associated with marked decreases in susceptibility only to elvitegravir.

Table 4. Integrase inhibitor resistance mutations*.

66 92 121 138 140 143 147 148 153 155 157 263

T E F E G Y S Q S N E R
Raltegravir Q Y AK AS CHR G HRK HS Q
Elvitegravir I Q Y AK AS n/a G HRK Y HS Q K
*

The first row of letters contains the consensus amino acid at the position indicated by the number in the preceding row. All amino acids are indicated by their one letter code. INI-resistance mutations selected in patients receiving raltegravir251,257 or elvitegravir252 and characterized in vitro for susceptibility250,252,253,255. Mutations in bold are associated with > 5-10 fold decreased susceptibility256.

Other mutations selected in vitro or in vivo by raltegravir include the non-polymorphic mutations H183P, Y226DFH, S230R, and D232N, and the polymorphic mutations L74M, T97A, V151I, G163R, I203M, and S230N256.

Other mutations selected in vitro or in vivo by elvitegravir include the non-polymorphic mutation H51Y, Q95K, and Q146P. Additional integrase mutations selected by other investigational integrase inhibitors include the non-polymorphic mutations T125K, A128T, Q146K, N155S, K160D and the polymorphic mutations V72I, M154I, V165I and V201I249.

Fusion inhibitors

Enfuvirtide, the only licensed fusion inhibitor, inhibits the interaction of the heptad repeat (HR) 1 and 2 domains of gp41 by mimicking a part of HR2 (amino acids 127-162) that binds to a conserved part of HR1. It has antiviral activity approaching that of the most active ARV such as efavirenz, lopinavir/r, and raltegravir. However, resistance may develop rapidly in patients receiving enfuvirtide for salvage therapy who do not receive a sufficient number of additional active drugs. Indeed, the emergence of resistance strains followed by virologic rebound has been observed in some patients within two to four weeks258,259.

The extra-viral portion of gp41 is the most conserved region in the HIV-1 envelope glycoprotein and there is little naturally occurring variation in the HR1 binding site among the different group M subtypes260-265. Nonetheless, there is about 10-fold variation in enfuvirtide susceptibility among isolates from enfuvirtide-naive persons, possibly resulting from gp41 polymorphisms outside of the HR1 binding site266-269. However, despite the wider range in baseline enfuvirtide susceptibility than for other ARV, there is no evidence that enfuvirtide-naive patients infected with viruses at the lower ranges of enfuvirtide susceptibility respond less well to enfuvirtide267,269.

Mutations in gp41 codons 36 to 45, the region to which enfuvirtide binds, are primarily responsible for enfuvirtide resistance269-274. Table 5 lists the most commonly observed enfuvirtide resistance mutations in this region. A single mutation is generally associated with about 10-fold decreased susceptibility, whereas double mutations can decrease susceptibility more than 100-fold. Several accessory mutations in the HR2 region corresponding to the peptide sequence of enfuvirtide including N126K, N137K, and S138A appear to improve fitness in combination with specific mutations at positions 36-45274-277. Similar enfuvirtide resistance mutations appear to emerge in subtype B and non-B isolates278,279.

Table 5. Fusion inhibitor resistance mutations*.

G36 I37 V38 Q40 N42 N43 L44 L45
Enfuvirtide DEVS V EAMG H T DKS M M
*

Mutations in bold reduce enfuvirtide susceptibility > 10-fold in site-directed mutants and most clinical isolates. N42S is the only common polymorphism between codons 36 to 45. It occurs in about 15% of untreated isolates and does not decrease enfuvirtide susceptibility269. Most other mutations at these positions are likely to have been selected by enfuvirtide, although their effect on enfuvirtide susceptibility may not have been reported. Several accessory mutations in the HR2 region corresponding to the peptide sequence of enfuvirtide including N126K, N137K, and S138A have been shown to emerge to improve fitness in combination with specific mutations at positions 36 to 45274-277.

Enfuvirtide-resistant HIV-1 isolates replicate less well than enfuvirtide-susceptible isolates, as evidenced by in vitro competition studies280 and by the rapid reversion to wild-type that occurs in patients who discontinue enfuvirtide281. There are some conflicting data on the clinical benefit of continued therapy in the presence of incomplete virologic suppression. One study showed that interruption of therapy was associated with a mean increase in plasma HIV-1 RNA levels of just 0.2 log10 and no decrease in CD4 count281. However, other studies have suggested that some enfuvirtide resistance mutations, particularly those at position 38, may be associated with CD4 count increases282, possibly because mutations at this position may decrease virus replication or render the virus more susceptible to neutralizing antibodies that target fusion intermediates283.

CCR5 inhibitors

The licensed small molecule inhibitor maraviroc and the investigational small molecule inhibitor vicriviroc (formerly SCH-D) allosterically inhibit HIV-1 gp120 binding to the seven-transmembrane G protein-coupled CCR5 coreceptor. Whereas HIV-1 gp120 binds to the N-terminus and second extracellular loop region of CCR5284, site-directed mutagenesis and molecular modeling studies suggest that most small molecule inhibitors bind to a pocket formed by the transmembrane helices285-290.

The HIV-1 gp120 has a highly variable sequence, and different HIV-1 isolates display variable susceptibility to inhibition by different ligands and small molecule inhibitors291,292. Nonetheless, there appear to be minimal differences in the susceptibility of wild-type viruses (even those belonging to different subtypes), to maraviroc293 and vicriviroc294, suggesting that these inhibitors disrupt a highly conserved protein-protein interaction. Moreover, in vitro passage experiments have generally demonstrated that high-level resistance emerges only after several months of passage, suggesting that the genetic barrier to resistance to CCR5 inhibitors is not low295-298. Nonetheless, HIV-1 may escape from CCR5 inhibition by developing CCR5 inhibitor resistance or by utilizing the CXCR4 coreceptor.

CCR5 inhibitor resistance

The genotypic and phenotypic mechanisms of CCR5 inhibitor resistance and cross-resistance are complex and poorly understood. Dose-response curves suggest that there are at least two phenotypic mechanisms of resistance: (i) enhanced binding to unbound CCR5, manifested by a shift in the typical sigmoid dose-response curves resulting from the requirement of a several-fold increase in the concentration of inhibitor required to suppress virus in a manner similar to wild-type292,297-300; and (ii) enhanced binding to a CCR5-inhibitor complex, manifested by a plateau in the maximal percent inhibition (MPI) regardless of inhibitor concentration. Such plateaus have been observed when testing both individual virus clones as well as virus populations within a clinical isolate. For a virus clone, the level of the MPI plateau is expected to be inversely proportional to the relative affinity of the clone for the bound compared with the unbound form of the receptor298. For a virus isolate, the level of the MPI plateau is expected to be inversely proportional to the number of viruses within the isolate with affinity for the bound compared with the unbound form of the receptor292.

Viruses with high levels of CCR5 inhibitor resistance (> 1,000-fold reductions in IC50 as well as an MPI plateau) have been identified during in vitro passage experiments with most CCR5 inhibitors297-299,301,302. The amino acid changes responsible for resistance may be entirely within the V3 loop298,299, entirely outside of the V3 loop297, or may result from synergistic interactions between substitutions in the V3 loop and other parts of env302. These amino acid changes may include known polymorphisms as well as novel substitutions, insertions, and deletions. Further complicating the genetic basis of CCR5 inhibitor resistance is the observation that the same inhibitor may select for different mutations in different virus isolates297,298,302.

The mechanisms of CCR5 inhibitor resistance in vivo may be even more complicated than those that have been observed to emerge in vitro. First, virus isolates from the majority of patients developing virologic failure while receiving maraviroc303 or vicriviroc304 have not demonstrated phenotypic resistance. Second, the few viruses with phenotypic resistance (four of 37 for maraviroc and one of seven for vicriviroc) have demonstrated only subtle MPI reductions rather than the MPI reductions and large increases in IC50 that have been observed during the emergence of resistance in vitro. Finally, the mutations that have been observed in vivo have been highly variable, differing for each virus isolate303,304.

CXCR4 tropism

At the time of initial HIV-1 infection, at least 80-90% of patients have viruses that exclusively use CCR5 as their coreceptor (R5 tropic). During the course of infection, about 50% of patients with subtype B infections are eventually found to harbor viruses that use the CXCR4 coreceptor (X4 tropic)305-308. The emergence of X4 tropism usually occurs in later disease stages and, in the absence of ARV therapy, is followed by accelerated CD4 cell depletion. When X4-tropic viruses emerge, they usually co-circulate with R5-tropic viruses as minor variants306,309-311. Some X4-tropic viruses are also R5 tropic, although most such dual-tropic HIV-1 clones usually infect only one of the two coreceptors efficiently311,312.

The main determinants for coreceptor tropism are in the V3 loop, although changes outside of the V3 loop may also influence tropism, either in combination or independently of V3 changes284,313-319. The presence of positively charged amino acids at positions 11 and 25 in the V3 loop, combined with several other V3 sequence characteristics, have a specificity of about 90% and sensitivity of 70-80% for predicting X4 tropism of individual virus clones belonging to subtype B284,316-318. However, the number and type of mutations by which an R5-tropic virus becomes X4 tropic is complex and depends on the sequence of the baseline R5 virus284,313-315,320. Preliminary data also suggests that the frequency and genetic basis for tropism switches may be different for different subtypes284,312.

A phenotypic assay has recently been developed to assess the tropism of complete env genes (gp120 plus gp41) amplified from patient samples (Trofile™, Monogram Biosciences)311. Amplified env genes are ligated into env expression test vectors (eETVs), which following co-transfection with env-deleted genomic vectors are used to create a population of pseudovirions. CD4+/U87 cells expressing CCR5 or CXCR4 are inoculated with these pseudovirions, and infection of each cell type is measured using a luciferase-based reporter system. In reconstruction experiments, X4-tropic variants can also be detected even when they constitute 1-5% of a mixed virus population311,321. However, because the amplification sensitivity of the assay is reliable only when plasma HIV-1 RNA levels are > 1,000 copies/ml, the full sensitivity of the assay will be achievable only in patients with plasma HIV-1 RNA levels > 10,000 copies/ml.

The factors responsible for the emergence of X4 tropism and for the proportion of X4 variants relative to R5 variants in those patients in whom X4-tropic viruses do emerge are not known. Yet, these factors have implications for detecting X4-tropic viruses to determine whether a CCR5 inhibitor will be effective. In a 10-day monotherapy study of maraviroc in 62 patients with CD4 counts > 250 cells/ul and the absence of X4 variants by Trofile™ testing, X4 emergence and virologic rebound occurred in two patients322,323. Phylogenetic analysis of env clones from pre- and posttreatment time points indicated that the X4 variants probably emerged by outgrowth from a pretreatment X4 reservoir323. Considering that seven patients had been excluded from this study owing to the presence of X4 variants at screening, X4 variants were detected successfully at baseline in only seven of nine cases322. Similar findings were reported in patients receiving maraviroc in the MOTIVATE I and II trials324, as well as in clinical trials of vicriviroc325 and aplaviroc326.

Improved sensitivity for detecting X4 variants is required to ensure that CCR5 inhibitors are optimally used. The Trofile™ test is more sensitive than genotypic methods for detecting X4-tropic viruses in clinical samples for at least two reasons. First, because the assay uses complete patient-derived env genes, it can detect X4 tropism even when the changes responsible are outside of the V3 loop327. Second, the assay is more sensitive at detecting X4-tropic variants that are below the 20-30% limit of detection of standard population-based genotypic assays. Indeed, this factor alone appears to be responsible for the drop in sensitivity for V3 genotyping from 70-80% on individual clones to 30% on clinical samples318,327. Novel genotypic approaches such as ultra-deep pyrosequencing methods that simultaneously sequence multiple individual clones in a patient sample328,329 and novel bioinformatic approaches for analyzing these sets of sequences will be required to attain sensitivities approaching that of phenotypic assays. Although the complex genetic basis for coreceptor tropism poses a hurdle for genotypic relative to phenotypic approaches, this drawback may be offset if genotypic methods are capable of identifying transitional R5 to X4 variants that may be surrogates for the presence of low-level X4 emergence.

References

  • 1.Hammer S, Saag M, Schechter M, et al. Treatment for adult HIV infection: 2006 recommendations of the International AIDS Society-USA panel. JAMA. 2006;296:827–43. doi: 10.1001/jama.296.7.827. [DOI] [PubMed] [Google Scholar]
  • 2.Panel on Antiretroviral Guidelines for Adult and Adolescents. Guidelines for the use of antiretroviral agents in HIV-1-infected adults and adolescents. Departments of Health and Human Services; Jan 29, 2008. pp. 1–128. [Google Scholar]
  • 3.Deval J, Selmi B, Boretto J, et al. The molecular mechanism of multidrug resistance by the Q151M HIV-1 reverse transcriptase and its suppression using alpha-boranophosphate nucleotide analogs. J Biol Chem. 2002;277:42097–104. doi: 10.1074/jbc.M206725200. [DOI] [PubMed] [Google Scholar]
  • 4.Deval J, White K, Miller M, et al. Mechanistic basis for reduced viral and enzymatic fitness of HIV-1 reverse transcriptase containing both K65R and M184V mutations. J Biol Chem. 2004;279:509–16. doi: 10.1074/jbc.M308806200. [DOI] [PubMed] [Google Scholar]
  • 5.Arion D, Kaushik N, McCormick S, Borkow G, Parniak M. Phenotypic mechanism of HIV-1 resistance to 3′-azido-3′-deoxythymidine (AZT): increased polymerization processivity and enhanced sensitivity to pyro-phosphate of the mutant viral reverse transcriptase. Biochemistry. 1998;37:15908–17. doi: 10.1021/bi981200e. [DOI] [PubMed] [Google Scholar]
  • 6.Meyer P, Matsuura S, Mian A, So A, Scott W. A mechanism of AZT resistance: an increase in nucleotide-dependent primer unblocking by mutant HIV-1 reverse transcriptase. Mol Cell. 1999;4:35–43. doi: 10.1016/s1097-2765(00)80185-9. [DOI] [PubMed] [Google Scholar]
  • 7.Boyer P, Sarafianos S, Arnold E, Hughes S. Selective excision of AZTMP by drug-resistant HIV reverse transcriptase. J Virol. 2001;75:4832–42. doi: 10.1128/JVI.75.10.4832-4842.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Whitcomb J, Parkin N, Chappey C, Hellmann N, Petropoulos CJ. Broad NRTI cross-resistance in HIV-1 clinical isolates. J Infect Dis. 2003;188:992–1000. doi: 10.1086/378281. [DOI] [PubMed] [Google Scholar]
  • 9.Eron J. The treatment of antiretroviral-naive subjects with the 3TC/zidovudine combination: a review of North American (NUCA 3001) and European (NUCB 3001) trials. AIDS. 1996;10(Suppl 5):S11–19. doi: 10.1097/00002030-199612005-00003. [DOI] [PubMed] [Google Scholar]
  • 10.Ingrand D, Weber J, Boucher C, et al. Phase I/II study of 3TC (lamivudine) in HIV-positive, asymptomatic or mild AIDS-related complex patients: sustained reduction in viral markers. The Lamivudine European HIV Working Group. AIDS. 1995;9:1323–9. doi: 10.1097/00002030-199512000-00004. [DOI] [PubMed] [Google Scholar]
  • 11.Pluda J, Cooley T, Montaner J, et al. A phase I/II study of 2′-deoxy-3′-thiacytidine (lamivudine) in patients with advanced HIV infection. J Infect Dis. 1995;171:1438–47. doi: 10.1093/infdis/171.6.1438. [DOI] [PubMed] [Google Scholar]
  • 12.Miller V, Stark T, Loeliger A, Lange J. The impact of the M184V substitution in HIV-1 reverse transcriptase on treatment response. HIV Med. 2002;3:135–45. doi: 10.1046/j.1468-1293.2002.00101.x. [DOI] [PubMed] [Google Scholar]
  • 13.Diallo K, Gotte M, Wainberg M. Molecular impact of the M184V mutation in HIV-1 reverse transcriptase. Antimicrob Agents Chemother. 2003;47:3377–83. doi: 10.1128/AAC.47.11.3377-3383.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Vray M, Meynard J, Dalban C, et al. Predictors of the virologic response to a change in the antiretroviral treatment regimen in HIV-1-infected patients enrolled in a randomized trial comparing genotyping, phenotype and standard of care (Narval trial, ANRS 088) Antivir Ther. 2003;8:427–34. doi: 10.1177/135965350300800510. [DOI] [PubMed] [Google Scholar]
  • 15.Petropoulos C, Parkin N, Limoli K, et al. A novel phenotypic drug susceptibility assay for HIV-1. Antimicrob Agents Chemother. 2000;44:920–8. doi: 10.1128/aac.44.4.920-928.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Rhee S, Liu T, Ravela J, Gonzales M, Shafer R. Distribution of HIV-1 protease and reverse transcriptase mutation patterns in 4,183 persons undergoing genotypic resistance testing. Antimicrob Agents Chemother. 2004;48:3122–6. doi: 10.1128/AAC.48.8.3122-3126.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Brun-Vezinet F, Descamps D, Ruffault A, et al. Clinically relevant interpretation of genotype for resistance to abacavir. AIDS. 2003;17:1795–802. doi: 10.1097/00002030-200308150-00008. [DOI] [PubMed] [Google Scholar]
  • 18.Winters M, Bosch R, Albrecht M, Katzenstein D. Clinical impact of the M184V mutation on switching to didanosine or maintaining lamivudine treatment in NRTI-experienced patients. J Infect Dis. 2003;188:537–40. doi: 10.1086/377742. [DOI] [PubMed] [Google Scholar]
  • 19.Lanier E, Ait-Khaled M, Scott J, et al. Antiviral efficacy of abacavir in antiretroviral therapy-experienced adults harboring HIV-1 with specific patterns of resistance to NRTI. Antivir Ther. 2004;9:37–45. doi: 10.1177/135965350400900102. [DOI] [PubMed] [Google Scholar]
  • 20.Marcelin A, Flandre P, Pavie J, et al. Clinically relevant genotype interpretation of resistance to didanosine. Antimicrob Agents Chemother. 2005;49:1739–44. doi: 10.1128/AAC.49.5.1739-1744.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Molina J, Marcelin A, Pavie J, et al. Didanosine in HIV-1-infected patients experiencing failure of antiretroviral therapy: a randomized placebo-controlled trial. J Infect Dis. 2005;191:8407. doi: 10.1086/428094. [DOI] [PubMed] [Google Scholar]
  • 22.Eron J, Bosch R, Bettendorf D, Petch L, Fiscus S, Frank I. The effect of lamivudine therapy and M184V on the antiretroviral activity of didanosine. J Acquir Immune Defic Syndr. 2007;45:249–51. doi: 10.1097/QAI.0b013e318050d61f. [DOI] [PubMed] [Google Scholar]
  • 23.Vermeiren H, Van Craenenbroeck E, Alen P, Bacheler L, Picchio G, Lecocq P. Prediction of HIV-1 drug susceptibility phenotype from the viral genotype using linear regression modeling. J Virol Methods. 2007;145:47–55. doi: 10.1016/j.jviromet.2007.05.009. [DOI] [PubMed] [Google Scholar]
  • 24.Bennett D, McCormick L, Kline R, et al. U.S. surveillance of HIV drug resistance at diagnosis using HIV diagnostic sera. 12th CROI; Boston, USA. 2005. Abstract 674. [Google Scholar]
  • 25.Fessel W, Rhee S, Klein D, et al. Low risk of initial antiretroviral treatment failure in patients with wild-type HIV-1 by standard genotypic resistance testing. 15th CROI; Boston, USA. 2008. abstract 892. [Google Scholar]
  • 26.Yahi N, Tamalet C, Tourres C, et al. Mutation patterns of the reverse transcriptase and protease genes in HIV-1-infected patients undergoing combination therapy: survey of 787 sequences. J Clin Microbiol. 1999;37:4099–106. doi: 10.1128/jcm.37.12.4099-4106.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Gonzales M, Wu T, Taylor J, et al. Extended spectrum of HIV-1 reverse transcriptase mutations in patients receiving multiple nucleoside analog inhibitors. AIDS. 2003;17:791–9. doi: 10.1097/01.aids.0000050860.71999.23. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Marcelin A, Delaugerre C, Wirden M, et al. Thymidine analog reverse transcriptase inhibitors resistance mutations profiles and association to other NRTI resistance mutations observed in the context of virologic failure. J Med Virol. 2004;72:162–5. doi: 10.1002/jmv.10550. [DOI] [PubMed] [Google Scholar]
  • 29.Miller M, Margot N, Lu B, et al. Genotypic and phenotypic predictors of the magnitude of response to tenofovir disoproxil fumarate treatment in antiretroviral-experienced patients. J Infect Dis. 2004;189:837–46. doi: 10.1086/381784. [DOI] [PubMed] [Google Scholar]
  • 30.Cozzi-Lepri A, Ruiz L, Loveday C, et al. Thymidine analog mutation profiles: factors associated with acquiring specific profiles and their impact on the virologic response to therapy. Antivir Ther. 2005;10:791–802. [PubMed] [Google Scholar]
  • 31.De Luca A, Di Giambenedetto S, Romano L, et al. Frequency and treatment-related predictors of thymidine-analog mutation patterns in HIV-1 isolates after unsuccessful antiretroviral therapy. J Infect Dis. 2006;193:1219–22. doi: 10.1086/502976. [DOI] [PubMed] [Google Scholar]
  • 32.Rhee S, Liu T, Holmes S, Shafer R. HIV-1 subtype B protease and reverse transcriptase amino acid covariation. PLoS Comput Biol. 2007;3:e87. doi: 10.1371/journal.pcbi.0030087. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Yahi N, Tamalet C, Tourres C, Tivoli N, Fantini J. Mutation L210W of HIV-1 reverse transcriptase in patients receiving combination therapy; incidence, association with other mutations, and effects on the structure of mutated reverse transcriptase. J Biomed Sci. 2000;7:507–13. doi: 10.1007/BF02253366. [DOI] [PubMed] [Google Scholar]
  • 34.Marcelin A, Flandre P, Furco A, Wirden M, Molina J, Calvez V. Impact of HIV-1 reverse transcriptase polymorphism at codons 211 and 228 on virologic response to didanosine. Antivir Ther. 2006;11:693–9. [PubMed] [Google Scholar]
  • 35.De Luca A, Giambenedetto S, Trotta M, et al. Improved interpretation of genotypic changes in the HIV-1 reverse transcriptase coding region that determine the virologic response to didanosine. J Infect Dis. 2007;196:1645–53. doi: 10.1086/522231. [DOI] [PubMed] [Google Scholar]
  • 36.Yerly S, Rakik A, De Loes SK, et al. Switch to unusual amino acids at codon 215 of the HIV-1 reverse transcriptase gene in seroconvertors infected with zidovudine-resistant variants. J Virol. 1998;72:3520–3. doi: 10.1128/jvi.72.5.3520-3523.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Garcia-Lerma J, Nidtha S, Blumoff K, Weinstock H, Heneine W. Increased ability for selection of zidovudine resistance in a distinct class of wild-type HIV-1 from drug-naive persons. Proc Natl Acad Sci USA. 2001;98:13907–12. doi: 10.1073/pnas.241300698. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Delaugerre C, Roudiere L, Peytavin G, Rouzioux C, Viard J, Chaix M. Selection of a rare resistance profile in an HIV-1-infected patient exhibiting a failure to an antiretroviral regimen including tenofovir DF. J Clin Virol. 2005;32:241–4. doi: 10.1016/j.jcv.2004.05.020. [DOI] [PubMed] [Google Scholar]
  • 39.Sluis-Cremer N, Sheen C, Zelina S, Argoti Torres P, Parikh U, Mellors J. Molecular mechanism by which K70E in HIV-1 reverse transcriptase confers resistance to NRTI. Antimicrob Agents Chemother. 2006;51:48–53. doi: 10.1128/AAC.00683-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Bradshaw D, Malik S, Booth C, et al. Novel drug resistance pattern associated with the mutations K70G and M184V in HIV-1 reverse transcriptase. Antimicrob Agents Chemother. 2007;51:4489–91. doi: 10.1128/AAC.00687-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Rhee S, Fessel W, Zolopa A, et al. HIV-1 protease and reverse-transcriptase mutations: correlations with antiretroviral therapy in subtype B isolates and implications for drug-resistance surveillance. J Infect Dis. 2005;192:456–65. doi: 10.1086/431601. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Delaugerre C, Mouroux M, Yvon-Groussin A, et al. Prevalence and conditions of selection of E44D/A and V118I HIV-1 reverse transcriptase mutations in clinical practice. Antimicrob Agents Chemother. 2001;45:946–8. doi: 10.1128/AAC.45.3.946-948.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Montes B, Segondy M. Prevalence of the mutational pattern E44D/A and/or V118I in the reverse transcriptase gene of HIV-1 in relation to treatment with nucleoside analog RT inhibitors. J Med Virol. 2002;66:299–303. doi: 10.1002/jmv.2145. [DOI] [PubMed] [Google Scholar]
  • 44.Hertogs K, Bloor S, De Vroey V, et al. A novel HIV-1 reverse transcriptase mutational pattern confers phenotypic lamivudine resistance in the absence of mutation 184V. Antimicrob Agents Chemother. 2000;44:568–73. doi: 10.1128/aac.44.3.568-573.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Perno C, Cozzi-Lepri A, Balotta C, et al. Impact of mutations conferring reduced susceptibility to lamivudine on the response to antiretroviral therapy. Antivir Ther. 2001;6:195–8. [PubMed] [Google Scholar]
  • 46.Walter H, Schmidt B, Werwein M, Schwingel E, Korn K. Prediction of abacavir resistance from genotypic data: impact of zidovudine and lamivudine resistance in vitro and in vivo. Antimicrob Agents Chemother. 2002;46:89–94. doi: 10.1128/AAC.46.1.89-94.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Romano L, Venturi G, Bloor S, et al. Broad nucleoside-analog resistance implications for HIV-1 reverse-transcriptase mutations at codons 44 and 118. J Infect Dis. 2002;185:898–904. doi: 10.1086/339706. [DOI] [PubMed] [Google Scholar]
  • 48.Stoeckli T, MaWhinney S, Uy J, et al. Phenotypic and genotypic analysis of biologically cloned HIV-1 isolates from patients treated with zidovudine and lamivudine. Antimicrob Agents Chemother. 2002;46:4000–3. doi: 10.1128/AAC.46.12.4000-4003.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Girouard M, Diallo K, Marchand B, McCormick S, Gotte M. Mutations E44D and V118I in the reverse transcriptase of HIV-1 play distinct mechanistic roles in dual resistance to AZT and 3TC. J Biol Chem. 2003;278:34403–10. doi: 10.1074/jbc.M303528200. [DOI] [PubMed] [Google Scholar]
  • 50.Saberg P, Koppel K, Bratt G, et al. The reverse transcriptase mutation V118I is associated with virologic failure on abacavir-based ART in HIV-1 infection. Scand J Infect Dis. 2004;36:40–5. doi: 10.1080/00365540310017249. [DOI] [PubMed] [Google Scholar]
  • 51.Gianotti N, Galli L, Boeri E, et al. The 118I reverse transcriptase mutation is the only independent genotypic predictor of virologic failure to a stavudine-containing salvage therapy in HIV-1-infected patients. J Acquir Immune Defic Syndr. 2006;41:447–52. doi: 10.1097/01.qai.0000209903.89878.80. [DOI] [PubMed] [Google Scholar]
  • 52.Zaccarelli M, Tozzi V, Lorenzini P, et al. The V118I mutation as a marker of advanced HIV infection and disease progression. Antivir Ther. 2007;12:163–8. [PubMed] [Google Scholar]
  • 53.Svicher V, Sing T, Santoro M, et al. Involvement of novel HIV-1 reverse transcriptase mutations in the regulation of resistance to nucleoside inhibitors. J Virol. 2006;80:7186–98. doi: 10.1128/JVI.02084-05. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Ceccherini-Silberstein F, Cozzi-Lepri A, Ruiz L, et al. Impact of HIV-1 reverse transcriptase polymorphism F214L on virologic response to thymidine analog-based regimens in ART-naive and ART-experienced patients. J Infect Dis. 2007;196:1180–90. doi: 10.1086/521678. [DOI] [PubMed] [Google Scholar]
  • 55.Moyle G, DeJesus E, Cahn P, et al. Abacavir once or twice daily combined with once-daily lamivudine and efavirenz for the treatment of antiretroviral-naive HIV-infected adults: results of the Ziagen Once Daily in Antiretroviral Combination Study. J Acquir Immune Defic Syndr. 2005;38:417–25. doi: 10.1097/01.qai.0000147521.34369.c9. [DOI] [PubMed] [Google Scholar]
  • 56.Gallant J, DeJesus E, Arribas J, et al. Tenofovir DF, emtricitabine, and efavirenz vs. zidovudine, lamivudine, and efavirenz for HIV. N Engl J Med. 2006;354:251–60. doi: 10.1056/NEJMoa051871. [DOI] [PubMed] [Google Scholar]
  • 57.Eron J, Yeni P, Gathe J, et al. The KLEAN study of fosamprenavir-ritonavir versus lopinavir-ritonavir, each in combination with abacavir-lamivudine, for initial treatment of HIV infection over 48 weeks: a randomised non-inferiority trial. Lancet. 2006;368:476–82. doi: 10.1016/S0140-6736(06)69155-1. [DOI] [PubMed] [Google Scholar]
  • 58.Lanier E, Givens N, Stone C, et al. Effect of concurrent zidovudine use on the resistance pathway selected by abacavir-containing regimens. HIV Med. 2004;5:394–9. doi: 10.1111/j.1468-1293.2004.00243.x. [DOI] [PubMed] [Google Scholar]
  • 59.Staszewski S, Dauer B, Locher L, Moesch M, Gute P, Stuermer M. Intensification of a failing regimen with zidovudine may cause sustained virologic suppression in the presence of the K65R Mutation. 13th CROI; 2006. abstract 635. [DOI] [PubMed] [Google Scholar]
  • 60.Antinori A, Trotta M, Lorenzini P, et al. Virologic response to salvage therapy in HIV-infected persons carrying the reverse transcriptase K65R mutation. Antivir Ther. 2007;12:1175–83. [PubMed] [Google Scholar]
  • 61.Rhee S, Taylor J, Wadhera G, Ben-Hur A, Brutlag D, Shafer R. Genotypic predictors of HIV-1 drug resistance. Proc Natl Acad Sci USA. 2006;103:17355–60. doi: 10.1073/pnas.0607274103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Wirden M, Roquebert B, Derache A, et al. Risk factors for selection of the L74I reverse transcriptase mutation in HIV-1-infected patients. Antimicrob Agents Chemother. 2006;50:2553–6. doi: 10.1128/AAC.00092-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Berkhout B, Back N, de Ronde A, et al. Identification of alternative amino acid substitutions in drug-resistant variants of the HIV-1 reverse transcriptase. AIDS. 2006;20:1515–20. doi: 10.1097/01.aids.0000237367.56864.75. [DOI] [PubMed] [Google Scholar]
  • 64.Gallant J, Staszewski S, Pozniak A, et al. Efficacy and safety of TDF vs. stavudine in combination therapy in antiretroviral-naive patients: a 3-year randomized trial. JAMA. 2004;292:191–201. doi: 10.1001/jama.292.2.191. [DOI] [PubMed] [Google Scholar]
  • 65.Margot N, Lu B, Cheng A, Miller M. Resistance development over 144 weeks in treatment-naive patients receiving TDF or stavudine with lamivudine and efavirenz in Study 903. HIV Med. 2006;7:442–50. doi: 10.1111/j.1468-1293.2006.00404.x. [DOI] [PubMed] [Google Scholar]
  • 66.Sosa N, Hill-Zabala C, Dejesus E, et al. Abacavir and lamivudine fixed-dose combination tablet once daily compared with abacavir and lamivudine twice daily in HIV-infected patients over 48 weeks ( ESS30008, SEAL) J Acquir Immune Defic Syndr. 2005;40:422–7. doi: 10.1097/01.qai.0000184859.24071.bd. [DOI] [PubMed] [Google Scholar]
  • 67.Smith K, Weinberg W, Dejesus E, et al. AIDS Res Ther. Vol. 5. 2008. Fosamprenavir or atazanavir once daily boosted with ritonavir 100 mg, plus tenofovir/emtricitabine, for the initial treatment of HIV infection: 48-week results of ALERT; p. 5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Saag M, Cahn P, Raffi F, et al. Efficacy and safety of emtricitabine vs. stavudine in combination therapy in antiretroviral-naive patients: a randomized trial. JAMA. 2004;292:180–9. doi: 10.1001/jama.292.2.180. [DOI] [PubMed] [Google Scholar]
  • 69.Descamps D, Ait-Khalid M, Craig C, et al. Rare selection of the K65R mutation in antiretroviral naive patients failing a first-line abacavir/lamivudine-containing HAART regimen. Antivir Ther. 2006;11:701–5. [PubMed] [Google Scholar]
  • 70.Svarovskaia E, Margot N, Bae A, et al. Low-level K65R mutation in HIV-1 reverse transcriptase of treatment-experienced patients exposed to abacavir or didanosine. J Acquir Immune Defic Syndr. 2007;46:174–80. doi: 10.1097/QAI.0b013e31814258c0. [DOI] [PubMed] [Google Scholar]
  • 71.Parikh U, Bacheler L, Koontz D, Mellors J. The K65R mutation in HIV-1 reverse transcriptase exhibits bidirectional phenotypic antagonism with thymidine analog mutations. J Virol. 2006;80:4971–7. doi: 10.1128/JVI.80.10.4971-4977.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Parikh U, Zelina S, Sluis-Cremer N, Mellors J. Molecular mechanisms of bidirectional antagonism between K65R and thymidine analog mutations in HIV-1 reverse transcriptase. AIDS. 2007;21:1405–14. doi: 10.1097/QAD.0b013e3281ac229b. [DOI] [PubMed] [Google Scholar]
  • 73.Parikh U, Barnas D, Faruki H, Mellors J. Antagonism between the HIV-1 reverse transcriptase mutation K65R and thymidine-analog mutations at the genomic level. J Infect Dis. 2006;194:651–60. doi: 10.1086/505711. [DOI] [PubMed] [Google Scholar]
  • 74.Shafer R, Kozal M, Winters M, et al. Combination therapy with zidovudine and didanosine selects for drug-resistant HIV-1 strains with unique patterns of pol gene mutations. J Infect Dis. 1994;169:722–9. doi: 10.1093/infdis/169.4.722. [DOI] [PubMed] [Google Scholar]
  • 75.Roge B, Katzenstein T, Obel N, et al. K65R with and without S68: a new resistance profile in vivo detected in most patients failing abacavir, didanosine and stavudine. Antivir Ther. 2003;8:173–82. [PubMed] [Google Scholar]
  • 76.Shafer R, Smeaton LM, Robbins G, et al. Comparison of four-drug regimens and pairs of sequential three-drug regimens as initial therapy for HIV-1 infection. N Engl J Med. 2003;349:2304–15. doi: 10.1056/NEJMoa030265. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Rey D, Krebs M, Partisani M, et al. Virologic response of zidovudine, lamivudine, and TDF combination in antiretroviral-naive HIV-1-Infected Patients. J Acquir Immune Defic Syndr. 2006;43:530–4. doi: 10.1097/01.qai.0000245885.74133.d9. [DOI] [PubMed] [Google Scholar]
  • 78.DART Virology Group and Trial Team. Virologic response to a triple nucleoside/nucleotide analog regimen over 48 weeks in HIV-1-infected adults in Africa. AIDS. 2006;20:1391–9. doi: 10.1097/01.aids.0000233572.59522.45. [DOI] [PubMed] [Google Scholar]
  • 79.Masquelier B, Neau D, Boucher S, et al. Antiretroviral efficacy and virologic profile of a zidovudine/lamivudine/TDF combination therapy in antiretroviral-naive patients. Antivir Ther. 2006;11:827–30. [PubMed] [Google Scholar]
  • 80.Elion R, Cohen C, Dejesus E, et al. Once-daily abacavir/lamivudine/zidovudine plus TDF for the treatment of HIV-1 infection in antiretroviralnaive subjects: a 48-week pilot study. HIV Clin Trials. 2006;7:324–33. doi: 10.1310/hct0706-324. [DOI] [PubMed] [Google Scholar]
  • 81.Sturmer M, Staszewski S, Doerr H. Quadruple nucleoside therapy with zidovudine, lamivudine, abacavir and tenofovir in the treatment of HIV. Antivir Ther. 2007;12:695–703. [PubMed] [Google Scholar]
  • 82.Ross L, Dretler R, Gerondelis P, Rouse E, Lim M, Lanier E. A rare HIV reverse transcriptase mutation, K65N, confers reduced susceptibility to TDF, lamivudine and didanosine. AIDS. 2006;20:787–9. doi: 10.1097/01.aids.0000216387.60481.0e. [DOI] [PubMed] [Google Scholar]
  • 83.Bartlett J, Johnson J, Herrera G, et al. Long-term results of initial therapy with abacavir and lamivudine combined with efavirenz, amprenavir/ritonavir, or stavudine. J Acquir Immune Defic Syndr. 2006;43:284–92. doi: 10.1097/01.qai.0000243092.40490.26. [DOI] [PubMed] [Google Scholar]
  • 84.Van Houtte M, Staes M, Geretti A, Patterry T, Bacheler L. NRTI resistance associated with the RT mutation K70E in HIV-1. Antivir Ther. 2006;11:S160. abstract 144. [Google Scholar]
  • 85.Ross L, Gerondelis P, Liao Q, et al. Selection of the HIV-1 reverse transcriptase mutation K70E in antiretroviral-naive subjects treated with tenofovir/abacavir/lamivudine therapy. Antivir Ther. 2005;10:S102. abstract 92. [Google Scholar]
  • 86.Tisdale M, Alnadaf T, Cousens D. Combination of mutations in HIV-1 reverse transcriptase required for resistance to the carbocyclic nucleoside 1592U89. Antimicrob Agents Chemother. 1997;41:1094–8. doi: 10.1128/aac.41.5.1094. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Margot N, Miller M. In vitro combination studies of tenofovir and other nucleoside analogs with ribavirin against HIV-1. Antivir Ther. 2005;10:343–8. [PubMed] [Google Scholar]
  • 88.Fitzgibbon J, Howell R, Haberzettl C, Sperber S, Gocke D, Dubin D. HIV-1 pol gene mutations which cause decreased susceptibility to 2′,3′dideoxycytidine. Antimicrob Agents Chemother. 1992;36:153–7. doi: 10.1128/aac.36.1.153. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Lacey S, Larder B. Novel mutation (V75T) in HIV-1 reverse transcriptase confers resistance to 2′,3′-didehydro-2′,3′-dideoxythymidine in cell culture. Antimicrob Agents Chemother. 1994;38:1428–32. doi: 10.1128/aac.38.6.1428. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Winters M, Merigan T. Variants other than aspartic acid at codon 69 of the HIV-1 reverse transcriptase gene affect susceptibility to nucleoside analogs. Antimicrob Agents Chemother. 2001;45:2276–9. doi: 10.1128/AAC.45.8.2276-2279.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Selmi B, Boretto J, Navarro J, et al. The valine-to-threonine 75 substitution in HIV-1 reverse transcriptase and its relation with stavudine resistance. J Biol Chem. 2001;276:13965–74. doi: 10.1074/jbc.M009837200. [DOI] [PubMed] [Google Scholar]
  • 92.Lennerstrand J, Stammers D, Larder B. Biochemical mechanism of HIV-1 reverse transcriptase resistance to stavudine. Antimicrob Agents Chemother. 2001;45:2144–6. doi: 10.1128/AAC.45.7.2144-2146.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Brenner B, Oliveira M, Doualla-Bell F, et al. HIV-1 subtype C viruses rapidly develop K65R resistance to tenofovir in cell culture. AIDS. 2006;20:F9–13. doi: 10.1097/01.aids.0000232228.88511.0b. [DOI] [PubMed] [Google Scholar]
  • 94.Doualla-Bell F, Avalos A, Brenner B, et al. High prevalence of the K65R mutation in HIV-1 subtype C isolates from infected patients in Botswana treated with didanosine-based regimens. Antimicrob Agents Chemother. 2006;50:4182–5. doi: 10.1128/AAC.00714-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Hawkins C, Chaplin B, Idoko J, et al. Clinical and genotypic findings in HIV-infected patients with the K65R mutation failing stavudine, lamivudine, and nevirapine first line antiretroviral therapy. Antivir Ther. 2007;12:S78. doi: 10.1097/QAI.0b013e3181b06125. abstract 69. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Rey D, Schmitt M, Hoizey G, et al. Early virologic non-response to once daily combination of lamivudine, tenofovir, and nevirapine n ART-naive HIV-infected patients: preliminary results of the DAUFIN study. 14th CROI; Los Angeles. 2007. abstract 503. [Google Scholar]
  • 97.Winters M, Coolley K, Girard Y, et al. A 6-basepair insert in the reverse transcriptase gene of HIV-1 confers resistance to multiple nucleoside inhibitors. J Clin Invest. 1998;102:1769–75. doi: 10.1172/JCI4948. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Van Vaerenbergh K, Van Laethem K, Albert J, et al. Prevalence and characteristics of multinucleoside-resistant HIV-1 among European patients receiving combinations of nucleoside analogs. Antimicrob Agents Chemother. 2000;44:2109–17. doi: 10.1128/aac.44.8.2109-2117.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Masquelier B, Race E, Tamalet C, et al. Genotypic and phenotypic resistance patterns of HIV-1 variants with insertions or deletions in the RT: multicenter study of patients treated with RT inhibitors. Antimicrob Agents Chemother. 2001;45:1836–42. doi: 10.1128/AAC.45.6.1836-1842.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.McColl D, Margot N, Wulfsohn M, Coakley D, Cheng A, Miller M. Patterns of resistance emerging in HIV-1 from antiretroviral-experienced patients undergoing intensification therapy with TDF. J Acquir Immune Defic Syndr. 2004;37:1340–50. doi: 10.1097/00126334-200411010-00002. [DOI] [PubMed] [Google Scholar]
  • 101.Prado J, Franco S, Matamoros T, et al. Relative replication fitness of multi-nucleoside analog-resistant HIV-1 strains bearing a dipeptide insertion in the fingers subdomain of the reverse transcriptase and mutations at codons 67 and 215. Virology. 2004;326:103–12. doi: 10.1016/j.virol.2004.06.006. [DOI] [PubMed] [Google Scholar]
  • 102.Shirasaka T, Kavlick M, Ueno T, et al. Emergence of HIV-1 variants with resistance to multiple dideoxynucleosides in patients receiving therapy with dideoxynucleosides. Proc Natl Acad Sci USA. 1995;92:2398–402. doi: 10.1073/pnas.92.6.2398. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Iversen A, Shafer R, Wehrly K, et al. Multidrug-resistant HIV-1 strains resulting from combination antiretroviral therapy. J Virol. 1996;70:1086–90. doi: 10.1128/jvi.70.2.1086-1090.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Shafer R, Iversen A, Winters M, Aguiniga E, Katzenstein D, Merigan T. Drug resistance and heterogeneous long-term virologic responses of HIV-1-infected subjects to zidovudine and didanosine combination therapy. The AIDS Clinical Trials Group 143 Virology Team. J Infect Dis. 1995;172:70–8. doi: 10.1093/infdis/172.1.70. [DOI] [PubMed] [Google Scholar]
  • 105.Kosalaraksa P, Kavlick M, Maroun V, Le R, Mitsuya H. Comparative fitness of multi-dideoxynucleoside-resistant HIV-1 in an In vitro competitive HIV-1 replication assay. J Virol. 1999;73:5356–63. doi: 10.1128/jvi.73.7.5356-5363.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Garcia-Lerma J, Gerrish P, Wright A, Qari S, Heneine W. Evidence of a role for the Q151L mutation and the viral background in development of multiple dideoxynucleoside-resistant HIV- 1. J Virol. 2000;74:9339–46. doi: 10.1128/jvi.74.20.9339-9346.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Dykes C, Demeter L. Clinical significance of HIV-1 replication fitness. Clin Microbiol Rev. 2007;20:550–78. doi: 10.1128/CMR.00017-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Rodes B, Holguin A, Soriano V, et al. Emergence of drug resistance mutations in HIV-2-infected subjects undergoing antiretroviral therapy. J Clin Microbiol. 2000;38:1370–4. doi: 10.1128/jcm.38.4.1370-1374.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Brandin E, Lindborg L, Gyllensten K, et al. pol gene sequence variation in Swedish HIV-2 patients failing antiretroviral therapy. AIDS Res Hum Retroviruses. 2003;19:543–50. doi: 10.1089/088922203322230905. [DOI] [PubMed] [Google Scholar]
  • 110.Gallego O, de Mendoza C, Labarga P, et al. Long-term outcome of HIV-infected patients with multinucleoside-resistant genotypes. HIV Clin Trials. 2003;4:372–81. doi: 10.1310/X618-KWKJ-WCTQ-LQ2L. [DOI] [PubMed] [Google Scholar]
  • 111.Zaccarelli M, Perno C, Forbici F, et al. Q151M-mediated multinucleoside resistance: prevalence, risk factors, and response to salvage therapy. Clin Infect Dis. 2004;38:433–7. doi: 10.1086/381097. [DOI] [PubMed] [Google Scholar]
  • 112.Saracino A, Monno L, Scudeller L, et al. Impact of unreported HIV-1 reverse transcriptase mutations on phenotypic resistance to nucleoside and nonnucleoside inhibitors. J Med Virol. 2006;78:9–17. doi: 10.1002/jmv.20500. [DOI] [PubMed] [Google Scholar]
  • 113.Paolucci S, Baldanti F, Maga G, et al. Gln145Met/Leu changes in HIV-1 reverse transcriptase confer resistance to nucleoside and nonnucleoside analogs and impair virus replication. Antimicrob Agents Chemother. 2004;48:4611–7. doi: 10.1128/AAC.48.12.4611-4617.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Paolucci S, Baldanti F, Tinelli M, Maga G, Gerna G. Detection of a new HIV-1 reverse transcriptase mutation (Q145M) conferring resistance to nucleoside and nonnucleoside inhibitors in a patient failing HAART. AIDS. 2003;17:924–7. doi: 10.1097/00002030-200304110-00022. [DOI] [PubMed] [Google Scholar]
  • 115.Smith R, Klarmann G, Stray K, et al. A new point mutation (P157S) in the reverse transcriptase of HIV-1 confers low-level resistance to (-)-beta2′,3′-dideoxy-3′-thiacytidine. Antimicrob Agents Chemother. 1999;43:2077–80. doi: 10.1128/aac.43.8.2077. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Picard V, Angelini E, Maillard A, et al. Comparison of genotypic and phenotypic resistance patterns of HIV-1 isolates from patients treated with stavudine and didanosine or zidovudine and lamivudine. J Infect Dis. 2001;184:781–4. doi: 10.1086/323088. [DOI] [PubMed] [Google Scholar]
  • 117.Nikolenko G, Palmer S, Maldarelli F, Mellors J, Coffin J, Pathak V. Mechanism for nucleoside analog-mediated abrogation of HIV-1 replication: balance between RNaseH activity and nucleotide excision. Proc Natl Acad Sci USA. 2005;102:2093–8. doi: 10.1073/pnas.0409823102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Yap S, Sheen C, Fahey J, et al. N348I in the connection domain of HIV-1 reverse transcriptase confers zidovudine and nevirapine resistance. PLoS Med. 2007;4:e335. doi: 10.1371/journal.pmed.0040335. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Kemp S, Shi C, Bloor S, Harrigan P, Mellors J, Larder B. A novel polymorphism at codon 333 of HIV-1 reverse transcriptase can facilitate dual resistance to zidovudine and L-2′,3′-dideoxy-3′-thiacytidine. J Virol. 1998;72:5093–8. doi: 10.1128/jvi.72.6.5093-5098.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Nikolenko G, Delviks-Frankenberry K, Palmer S, et al. Mutations in the connection domain of HIV-1 reverse transcriptase increase 3′-azido-3′-deoxythymidine resistance. Proc Natl Acad Sci USA. 2007;104:317–22. doi: 10.1073/pnas.0609642104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Brehm J, Koontz D, Meteer J, Pathak V, Sluis-Cremer N, Mellors J. Selection of mutations in the connection and RNaseH domains of HIV-1 reverse transcriptase that increase resistance to 3′-azido-3′-dideoxythymidine. J Virol. 2007;81:7852–9. doi: 10.1128/JVI.02203-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Zelina S, Sheen C, Radzio J, Mellors J, Sluis-Cremer N. Mechanisms by which the G333D mutation in HIV-1 reverse transcriptase facilitates dual resistance to zidovudine and lamivudine. Antimicrob Agents Chemother. 2008;52:157–63. doi: 10.1128/AAC.00904-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Delviks-Frankenberry K, Nikolenko G, Barr R, Pathak V. Mutations in HIV-1 RNaseH primer grip enhance 3′-azido-3′-deoxythymidine resistance. J Virol. 2007;81:6837–45. doi: 10.1128/JVI.02820-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Roquebert B, Wirden M, Simon A, et al. Relationship between mutations in HIV-1 RNaseH domain and NRTI resistance mutations in naive and pretreated HIV infected patients. J Med Virol. 2007;79:207–11. doi: 10.1002/jmv.20788. [DOI] [PubMed] [Google Scholar]
  • 125.Ntemgwa M, Wainberg M, Oliveira M, et al. Variations in reverse transcriptase and RNaseH domain mutations in HIV-1 clinical isolates are associated with divergent phenotypic resistance to zidovudine. Antimicrob Agents Chemother. 2007;51:3861–9. doi: 10.1128/AAC.00646-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Sarafianos S, Das K, Hughes S, Arnold E. Taking aim at a moving target: designing drugs to inhibit drug-resistant HIV-1 reverse transcriptases. Curr Opin Struct Biol. 2004;14:716–30. doi: 10.1016/j.sbi.2004.10.013. [DOI] [PubMed] [Google Scholar]
  • 127.Ren J, Stammers D. Structural basis for drug resistance mechanisms for nonnucleoside inhibitors of HIV reverse transcriptase. Virus Res. 2008;134:157–70. doi: 10.1016/j.virusres.2007.12.018. [DOI] [PubMed] [Google Scholar]
  • 128.Shulman N, Zolopa A, Passaro D, et al. Efavirenz-and adefovir dipivoxil-based salvage therapy in highly treatment-experienced patients: clinical and genotypic predictors of virologic response. J Acquir Immune Defic Syndr. 2000;23:221–6. doi: 10.1097/00126334-200003010-00002. [DOI] [PubMed] [Google Scholar]
  • 129.Lecossier D, Shulman N, Morand-Joubert L, et al. Detection of minority populations of HIV-1 expressing the K103N resistance mutation in patients failing nevirapine. J Acquir Immune Defic Syndr. 2005;38:37–42. doi: 10.1097/00126334-200501010-00007. [DOI] [PubMed] [Google Scholar]
  • 130.Huang W, Gamarnik A, Limoli K, Petropoulos C, Whitcomb J. Amino acid substitutions at position 190 of HIV-1 reverse transcriptase increase susceptibility to delavirdine and impair virus replication. J Virol. 2003;77:1512–23. doi: 10.1128/JVI.77.2.1512-1523.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Uhlmann E, Tebas P, Storch G, et al. Effects of the G190A substitution of HIV reverse transcriptase on phenotypic susceptibility of patient isolates to delavirdine. J Clin Virol. 2004;31:198–203. doi: 10.1016/j.jcv.2004.03.012. [DOI] [PubMed] [Google Scholar]
  • 132.Bacheler L, Jeffrey S, Hanna G, et al. Genotypic correlates of phenotypic resistance to efavirenz in virus isolates from patients failing NNRTI therapy. J Virol. 2001;75:4999–5008. doi: 10.1128/JVI.75.11.4999-5008.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Vingerhoets J, Azijn H, Fransen E, et al. TMC125 displays a high genetic barrier to the development of resistance: evidence from in vitro selection experiments. J Virol. 2005;79:12773–82. doi: 10.1128/JVI.79.20.12773-12782.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Antinori A, Zaccarelli M, Cingolani A, et al. Cross-resistance among NNRTI limits recycling efavirenz after nevirapine failure. AIDS Res Hum Retroviruses. 2002;18:835–8. doi: 10.1089/08892220260190308. [DOI] [PubMed] [Google Scholar]
  • 135.Delaugerre C, Rohban R, Simon A, et al. Resistance profile and cross-resistance of HIV-1 among patients failing a NNRTI-containing regimen. J Med Virol. 2001;65:445–8. [PubMed] [Google Scholar]
  • 136.Walmsley S, Kelly D, Tseng A, Humar A, Harrigan P. NNRTI failure impairs HIV-RNA responses to efavirenz-containing salvage antiretroviral therapy. AIDS. 2001;15:1581–4. doi: 10.1097/00002030-200108170-00019. [DOI] [PubMed] [Google Scholar]
  • 137.Madruga J, Cahn P, Grinsztejn B, 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. doi: 10.1016/S0140-6736(07)61047-2. [DOI] [PubMed] [Google Scholar]
  • 138.Lazzarin A, Campbell T, Clotet B, 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. doi: 10.1016/S0140-6736(07)61048-4. [DOI] [PubMed] [Google Scholar]
  • 139.Vingerhoets J, Buelens M, Peeters M, et al. Impact of baseline mutations on the virologic response to TMC125 in the phase III clinical trials DUET-1 and DUET-2. Antivir Ther. 2007;12:S34. abstract 32. [Google Scholar]
  • 140.Pelemans H, Esnouf R, Dunkler A, et al. Characteristics of the Pro225His mutation in HIV-1 reverse transcriptase that appears under selective pressure of dose-escalating quinoxaline treatment of HIV-1. J Virol. 1997;71:8195–203. doi: 10.1128/jvi.71.11.8195-8203.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Parkin N, Gupta S, Chappey C, Petropoulos C. The K101P and K103R/V179D mutations in HIV-1 reverse transcriptase confer resistance to NNRTI. Antimicrob Agents Chemother. 2006;50:351–4. doi: 10.1128/AAC.50.1.351-354.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Balzarini J, Pelemans H, Esnouf R, De Clercq E. A novel mutation (F227L) arises in the reverse transcriptase of HIV-1 on dose-escalating treatment of HIV type 1-infected cell cultures with the NNRTI thiocarboxanilide UC-781. AIDS Res Hum Retroviruses. 1998;14:255–60. doi: 10.1089/aid.1998.14.255. [DOI] [PubMed] [Google Scholar]
  • 143.Huang W, Parkin N, Lie Y, et al. A novel HIV-1 RT mutation (M230L) confers NNRTI resistance and dose-dependent stimulation of replication. Antivir Ther. 2000;5(Suppl 3):2425. [Google Scholar]
  • 144.Vingerhoets J, De Baere I, Azijin H, et al. Antiviral activity of TMC125 against a panel of site-directed mutants encompassing mutations observed in vitro and in vivo. 11th CROI; San Francisco. 2004. abstract 621. [Google Scholar]
  • 145.Su G, Yan J, Li Y, et al. In vitro selection and characterization of viruses resistant to R1206 a novel NNRTI. Antivir Ther. 2007;12:S35. abstract 33. [Google Scholar]
  • 146.Harrigan P, Salim M, Stammers DK, et al. A Mutation in the 3′ region of the HIV- 1 reverse transcriptase (Y318F) associated with NNRTI resistance. J Virol. 2002;76:6836–40. doi: 10.1128/JVI.76.13.6836-6840.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Rhee S, Gonzales M, Kantor R, Betts B, Ravela J, Shafer R. HIV reverse transcriptase and protease sequence database. Nucleic Acids Res. 2003;31:298–303. doi: 10.1093/nar/gkg100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Shafer R, Rhee S, Pillay D, et al. HIV-1 protease and reverse transcriptase mutations for drug resistance surveillance. AIDS. 2007;21:215–23. doi: 10.1097/QAD.0b013e328011e691. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Harrigan P, Hertogs K, Verbiest W, et al. Modest decreases in NNRTI susceptibility do not influence virologic outcome in patients receiving initial NNRTI-containing triple therapy. Antivir Ther. 2003;8:395–402. [PubMed] [Google Scholar]
  • 150.Bannister W, Ruiz L, Cozzi-Lepri A, et al. Comparison of genotypic resistance profiles and virologic response between patients starting nevirapine and efavirenz in EuroSIDA. AIDS. 2008;22:367–76. doi: 10.1097/QAD.0b013e3282f3cc35. [DOI] [PubMed] [Google Scholar]
  • 151.Brown A, Precious H, Whitcomb J, et al. Reduced susceptibility of HIV-1 from patients with primary HIV infection to NNRTI is associated with variation at novel amino acid sites. J Virol. 2000;74:10269–73. doi: 10.1128/jvi.74.22.10269-10273.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Ceccherini-Silberstein F, Svicher V, Sing T, et al. Characterization and structural analysis of novel mutations in HIV-1 reverse transcriptase involved in the regulation of resistance to nonnucleoside inhibitors. J Virol. 2007;81:11507–19. doi: 10.1128/JVI.00303-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Kleim J, Rosner M, Winkler I, et al. Selective pressure of a quinoxaline nonnucleoside inhibitor of HIV-1 reverse transcriptase on HIV-1 replication results in the emergence of NRTI-specific (RT Leu-74-->Val or Ile and Val-75-->Leu or Ile) HIV-1 mutants. Proc Natl Acad Sci USA. 1996;93:34–8. doi: 10.1073/pnas.93.1.34. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Koval C, Dykes C, Wang J, Demeter L. Relative replication fitness of efavirenz-resistant mutants of HIV-1: correlation with frequency during clinical therapy and evidence of compensation for the reduced fitness of K103N + L100I by the nucleoside resistance mutation L74V. Virology. 2006;353:184–92. doi: 10.1016/j.virol.2006.05.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Gao Y, Paxinos E, Galovich J, et al. Characterization of a subtype D HIV-1 isolate that was obtained from an untreated individual and that is highly resistant NNRTI. J Virol. 2004;78:5390–401. doi: 10.1128/JVI.78.10.5390-5401.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Eshleman S, Jones D, Galovich J, et al. Phenotypic drug resistance patterns in subtype A HIV-1 clones with NNRTI resistance mutations. AIDS Res Hum Retroviruses. 2006;22:289–93. doi: 10.1089/aid.2006.22.289. [DOI] [PubMed] [Google Scholar]
  • 157.Whitcomb J, Huang W, Limoli K, et al. Hypersusceptibility to NNRTI in HIV-1: clinical, phenotypic and genotypic correlates. AIDS. 2002;16:F41–7. doi: 10.1097/00002030-200210180-00002. [DOI] [PubMed] [Google Scholar]
  • 158.Shulman N, Bosch R, Mellors J, Albrecht M, Katzenstein D. Genetic correlates of efavirenz hypersusceptibility. AIDS. 2004;18:1781–5. doi: 10.1097/00002030-200409030-00006. [DOI] [PubMed] [Google Scholar]
  • 159.Condra J, Holder D, Schleif W, et al. Genetic correlates of in vivo viral resistance to indinavir, a HIV-1 protease inhibitor. J Virol. 1996;70:8270–6. doi: 10.1128/jvi.70.12.8270-8276.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Molla A, Korneyeva M, Gao Q, et al. Ordered accumulation of mutations in HIV protease confers resistance to ritonavir. Nat Med. 1996;2:760–6. doi: 10.1038/nm0796-760. [DOI] [PubMed] [Google Scholar]
  • 161.Kempf D, Isaacson J, King M, et al. Identification of genotypic changes in HIV protease that correlate with reduced susceptibility to the protease inhibitor lopinavir among viral isolates from protease inhibitor-experienced patients. J Virol. 2001;75:7462–9. doi: 10.1128/JVI.75.16.7462-7469.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Patick A, Mo H, Markowitz M, et al. Antiviral and resistance studies of AG1343, an orally bioavailable inhibitor of HIV protease. Antimicrob Agents Chemother. 1996;40:292–7. doi: 10.1128/aac.40.2.292. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Lawrence J, Schapiro J, Winters M, et al. Clinical resistance patterns and responses to two sequential protease inhibitor regimens in saquinavir and reverse transcriptase inhibitor-experienced persons. J Infect Dis. 1999;179:1356–64. doi: 10.1086/314751. [DOI] [PubMed] [Google Scholar]
  • 164.Walmsley S, Becker M, Zhang M, Humar A, Harrigan P. Predictors of virologic response in HIV-infected patients to salvage antiretroviral therapy that includes nelfinavir. Antivir Ther. 2001;6:47–54. [PubMed] [Google Scholar]
  • 165.Casado J, Dronda F, Hertogs K, et al. Efficacy, tolerance, and pharmacokinetics of the combination of stavudine, nevirapine, nelfinavir, and saquinavir as salvage regimen after ritonavir or indinavir failure. AIDS Res Hum Retroviruses. 2001;17:93–8. doi: 10.1089/08892220150217175. [DOI] [PubMed] [Google Scholar]
  • 166.Johnston E, Winters M, Rhee S, Merigan T, Schiffer C, Shafer R. Association of a novel HIV-1 protease substrate cleft mutation, L23I, with protease inhibitor therapy and in vitro drug resistance. Antimicrob Agents Chemother. 2004;48:4864–8. doi: 10.1128/AAC.48.12.4864-4868.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Winters B, Montaner J, Harrigan P, et al. Determination of clinically relevant cutoffs for HIV-1 phenotypic resistance estimates through a combined analysis of clinical trial and cohort data. J Acquir Immune Defic Syndr. 2008;48:26–34. doi: 10.1097/QAI.0b013e31816d9bf4. [DOI] [PubMed] [Google Scholar]
  • 168.Colonno R, Rose R, McLaren C, Thiry A, Parkin N, Friborg J. Identification of I50L as the signature ATV-resistance mutation in treatment-naive HIV-1-infected patients receiving ATV-containing regimens. J Infect Dis. 2004;189:1802–10. doi: 10.1086/386291. [DOI] [PubMed] [Google Scholar]
  • 169.Coakley E, Mass M, Parkin N. Atazanavir resistance in a PI-naïve patient treated with atazanavir/ritonavir associated with development of high-level atazanavir resistance and the N88S mutation in protease. 12th CROI; Boston. 2005. abstract 716. [Google Scholar]
  • 170.Vora S, Marcelin A, Gunthard H, et al. Clinical validation of atazanavir/ritonavir genotypic resistance score in protease inhibitor-experienced patients. AIDS. 2006;20:35–40. doi: 10.1097/01.aids.0000196179.11293.fc. [DOI] [PubMed] [Google Scholar]
  • 171.Pellegrin I, Breilh D, Ragnaud J, et al. Virologic responses to atazanavir-ritonavir-based regimens: resistance-substitutions score and pharmacokinetic parameters (Reyaphar study) Antivir Ther. 2006;11:421–9. [PubMed] [Google Scholar]
  • 172.Coakley E, Chappey C, Maa J, et al. Evaluation of phenotypic clinical cutoff for atazanavir/ritonavir in patients who changed only the protease inhibitor component of HAART: A confirmatory week 2 analysis of AI424-045. 13th CROI; Denver. 2006. abstract 634. [Google Scholar]
  • 173.Zolopa A, Towner W, Butcher D, Wang S, Maa J, Seekins D. Resistance profile after viral rebound on atazanavir-containing therapy: focus on protease inhibitor-naive subjects in the IMPACT study (BMS AI424-128) Antivir Ther. 2007;12:S86. abstract 77. [Google Scholar]
  • 174.Winters B, Montaner J, Harrigan R, et al. Development of VircoTYPE HIV-1 resistance analysis including clinical cut-offs for ritonavir-boosted atazanavir and fosamprenavir. Antivir Ther. 2007;12:S170. abstract 155. [Google Scholar]
  • 175.Zolopa A, Shafer R, Warford A, et al. HIV-1 genotypic resistance patterns predict response to saquinavir-ritonavir therapy in patients in whom previous PI therapy had failed. Ann Intern Med. 1999;131:813–21. doi: 10.7326/0003-4819-131-11-199912070-00003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Marcelin A, Dalban C, Peytavin G, et al. Clinically relevant interpretation of genotype and relationship to plasma drug concentrations for resistance to saquinavir-ritonavir in HIV-1 PI-experienced patients. Antimicrob Agents Chemother. 2004;48:4687–92. doi: 10.1128/AAC.48.12.4687-4692.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Marcelin A, Flandre P, de Mendoza C, et al. Clinical validation of saquinavir/ritonavir genotypic resistance score in PI-experienced patients. Antivir Ther. 2007;12:247. [PubMed] [Google Scholar]
  • 178.Dandache S, Sevigny G, Yelle J, et al. In vitro antiviral activity and cross-resistance profile of PL-100, a novel PI of HIV-1. Antimicrob Agents Chemother. 2007;51:4036–43. doi: 10.1128/AAC.00149-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Marcelin A, Flandre P, Molina J, et al. Genotypic analysis of the virologic response to fosamprenavir/ritonavir in clinical trials: CONTEXT and TRIAD. 14th CROI; Los Angeles. 2007. abstract 608. [Google Scholar]
  • 180.Pellegrin I, Breilh D, Coureau G, et al. Interpretation of genotype and pharmacokinetics for resistance to fosamprenavir-ritonavir-based regimens in antiretroviral-experienced patients. Antimicrob Agents Chemother. 2007;51:1473–80. doi: 10.1128/AAC.00481-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Masquelier B, Assoumou K, Descamps D, et al. Clinically validated mutation scores for HIV-1 resistance to fosamprenavir/ritonavir. J Antimicrob Chemother. 2008;61:1362–8. doi: 10.1093/jac/dkn127. [DOI] [PubMed] [Google Scholar]
  • 182.Friend J, Parkin N, Liegler T, Martin J, Deeks S. Isolated lopinavir resistance after virologic rebound of a ritonavir/lopinavir-based regimen. AIDS. 2004;18:1965–6. doi: 10.1097/00002030-200409240-00016. [DOI] [PubMed] [Google Scholar]
  • 183.Kagan R, Shenderovich M, Heseltine P, Ramnarayan K. Structural analysis of an HIV-1 protease I47A mutant resistant to the PI lopinavir. Protein Sci. 2005;14:1870–8. doi: 10.1110/ps.051347405. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Baxter J, Schapiro J, Boucher CA, et al. Genotypic changes in HIV-1 protease associated with reduced susceptibility and virologic response to the PI tipranavir. J Virol. 2006;80:10794–801. doi: 10.1128/JVI.00712-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Mo H, Parkin N, Stewart K, et al. Identification and structural characterization of I84C and I84A mutations that are associated with high-level resistance to HIV PI and impair viral replication. Antimicrob Agents Chemother. 2007;51:732–5. doi: 10.1128/AAC.00690-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Elston R, Scherer J, Hall D, et al. De-selection of the I50V mutation occurs in clinical isolates during aptivus/r (tipranavir/ritonavir)-based therapy. Antivir Ther. 2006;11:S102. abstract 92. [Google Scholar]
  • 187.Ziermann R, Limoli K, Das K, Arnold E, Petropoulos C, Parkin N. A mutation in HIV-1 protease, N88S, that causes in vitro hypersensitivity to amprenavir. J Virol. 2000;74:4414–9. doi: 10.1128/jvi.74.9.4414-4419.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Braun P, Walter H, Hoffman D, et al. Clinically relevant resensitization of PI saquinavir and atazanavir by L76V in multidrug-resistant HIV-1-infected patients. Antivir Ther. 2007;12:S142. abstract 129. [Google Scholar]
  • 189.Mammano F, Petit C, Clavel F. Resistance-associated loss of viral fitness in HIV-1: phenotypic analysis of protease and gag co-evolution in PI-treated patients. J Virol. 1998;72:7632–7. doi: 10.1128/jvi.72.9.7632-7637.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Nijhuis M, Schuurman R, de Jong D, et al. Increased fitness of drug resistant HIV-1 protease as a result of acquisition of compensatory mutations during suboptimal therapy. AIDS. 1999;13:2349–59. doi: 10.1097/00002030-199912030-00006. [DOI] [PubMed] [Google Scholar]
  • 191.Martinez-Picado J, Savara A, Sutton L, D'Aquila R. Replicative fitness of PI-resistant mutants of HIV-1. J Virol. 1999;73:3744–52. doi: 10.1128/jvi.73.5.3744-3752.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.van Maarseveen N, de Jong D, Boucher C, Nijhuis M. An increase in viral replicative capacity drives the evolution of PI-resistant HIV-1 in the absence of drugs. J Acquir Immune Defic Syndr. 2006;42:162–8. doi: 10.1097/01.qai.0000219787.65915.56. [DOI] [PubMed] [Google Scholar]
  • 193.Hoffman N, Schiffer C, Swanstrom R. Covariation of amino acid positions in HIV-1 protease. Virology. 2003;314:536–48. doi: 10.1016/s0042-6822(03)00484-7. [DOI] [PubMed] [Google Scholar]
  • 194.Perno C, Cozzi-Lepri A, Balotta C, et al. Secondary mutations in the protease region of HIV and virologic failure in drug-naive patients treated with PI-based therapy. J Infect Dis. 2001;184:983–91. doi: 10.1086/323604. [DOI] [PubMed] [Google Scholar]
  • 195.Perno C, Cozzi-Lepri A, Forbici F, et al. J Infect Dis. Vol. 189. 2004. Minor mutations in HIV protease at baseline and appearance of primary mutation 90M in patients for whom their first PI antiretroviral regimens failed; pp. 1983–7. [DOI] [PubMed] [Google Scholar]
  • 196.Wu T, Schiffer C, Gonzales M, et al. Mutation patterns and structural correlates in HIV-1 protease following different PI treatments. J Virol. 2003;77:4836–47. doi: 10.1128/JVI.77.8.4836-4847.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Parkin N, Chappey C, Petropoulos C. Improving lopinavir genotype algorithm through phenotype correlations: novel mutation patterns and amprenavir cross-resistance. AIDS. 2003;17:955–61. doi: 10.1097/00002030-200305020-00003. [DOI] [PubMed] [Google Scholar]
  • 198.Ceccherini-Silberstein F, Erba F, Gago F, et al. Identification of the minimal conserved structure of HIV-1 protease in the presence and absence of drug pressure. AIDS. 2004;18:F11–9. doi: 10.1097/01.aids.0000131394.76221.02. [DOI] [PubMed] [Google Scholar]
  • 199.Svicher V, Ceccherini-Silberstein F, Erba F, et al. Novel HIV-1 protease mutations potentially involved in resistance to PI. Antimicrob Agents Chemother. 2005;49:2015–25. doi: 10.1128/AAC.49.5.2015-2025.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Scherer J, Boucher C, Baxter J, Schapiro J, Kohlbrenner V, Hall D. Improving the prediction of virologic response to tipranavir: the development of a tipranavir weighted score. 11th European AIDS Conference; Madrid, Spain. 2007. abstract P3.4/07. [Google Scholar]
  • 201.De Meyer S, Vangeneugden T, van Baelen B, et al. Resistance profile of darunavir: combined 24-week results from the POWER trials. AIDS Res Hum Retroviruses. 2008;24:379–88. doi: 10.1089/aid.2007.0173. [DOI] [PubMed] [Google Scholar]
  • 202.De Meyer S, Dierynck I, Lathouwers E, et al. Identification of mutations predictive of a diminished response to darunavir/ritonavir: analysis of data from treatment-experienced patients in POWER 1, 2, 3, and DUET-1 and DUET-2. 6th European HIV Drug Resistance Workshop; Budapest, Hungary. 2008. abstract 54. [Google Scholar]
  • 203.Cote H, Brumme Z, Harrigan P. HIV-1 protease cleavage site mutations associated with PI cross-resistance selected by indinavir, ritonavir, and/or saquinavir. J Virol. 2001;75:589–94. doi: 10.1128/JVI.75.2.589-594.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204.Dauber D, Ziermann R, Parkin N, et al. Altered substrate specificity of drug-resistant HIV-1 protease. J Virol. 2002;76:1359–68. doi: 10.1128/JVI.76.3.1359-1368.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 205.Prabu-Jeyabalan M, Nalivaika E, King N, Schiffer C. Structural basis for co-evolution of a HIV-1 nucleocapsid-p1 cleavage site with a V82A drug-resistant mutation in viral protease. J Virol. 2004;78:12446–54. doi: 10.1128/JVI.78.22.12446-12454.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Malet I, Roquebert B, Dalban C, et al. Association of Gag cleavage sites to protease mutations and to virologic response in HIV-1 treated patients. J Infect. 2007;54:367–74. doi: 10.1016/j.jinf.2006.06.012. [DOI] [PubMed] [Google Scholar]
  • 207.Pettit S, Lindquist J, Kaplan A, Swanstrom R. Processing sites in the HIV-1 Gag-Pro-Pol precursor are cleaved by the viral protease at different rates. Retrovirology. 2005;2:66. doi: 10.1186/1742-4690-2-66. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208.Zhang Y, Imamichi H, Imamichi T, et al. Drug resistance during indinavir therapy is caused by mutations in the protease gene and in its Gag substrate cleavage sites. J Virol. 1997;71:6662–70. doi: 10.1128/jvi.71.9.6662-6670.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Verheyen J, Litau E, Sing T, et al. Compensatory mutations at the HIV cleavage sites p7/p1 and p1/p6-gag in therapy-naive and therapy-experienced patients. Antivir Ther. 2006;11:879–87. [PubMed] [Google Scholar]
  • 210.Kolli M, Lastere S, Schiffer CA. Co-evolution of nelfinavir-resistant HIV-1 protease and the p1-p6 substrate. Virology. 2006;347:405–9. doi: 10.1016/j.virol.2005.11.049. [DOI] [PubMed] [Google Scholar]
  • 211.Carrillo A, Stewart K, Sham H, et al. In vitro selection and characterization of HIV-1 variants with increased resistance to ABT-378, a novel. PI J Virol. 1998;72:7532–41. doi: 10.1128/jvi.72.9.7532-7541.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Maguire M, Guinea R, Griffin P, et al. Changes in HIV-1 Gag at positions L449 and P453 are linked to I50V protease mutants in vivo and cause reduction of sensitivity to amprenavir and improved viral fitness in vitro. J Virol. 2002;76:7398–406. doi: 10.1128/JVI.76.15.7398-7406.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213.Bally F, Martinez R, Peters S, Sudre P, Telenti A. Polymorphism of HIV-1 gag p7/p1 and p1/p6 cleavage sites: clinical significance and implications for resistance to PI. AIDS Res Hum Retroviruses. 2000;16:1209–13. doi: 10.1089/08892220050116970. [DOI] [PubMed] [Google Scholar]
  • 214.Nijhuis M, van Maarseveen N, Lastere S, et al. A novel substrate-based HIV-1 PI drug resistance mechanism. PLoS Med. 2007;4:e36. doi: 10.1371/journal.pmed.0040036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Kaufmann G, Suzuki K, Cunningham P, et al. Impact of HIV-1 protease, reverse transcriptase, cleavage site, and p6 mutations on the virologic response to quadruple therapy with saquinavir, ritonavir, and two nucleoside analogs. AIDS Res Hum Retroviruses. 2001;17:487–97. doi: 10.1089/08892220151126526. [DOI] [PubMed] [Google Scholar]
  • 216.Gatanaga H, Suzuki Y, Tsang H, et al. J Biol Chem. Vol. 277. 2002. Amino acid substitutions in Gag protein at noncleavage sites are indispensable for the development of a high multitude of HIV-1 resistance against PI; pp. 5952–61. [DOI] [PubMed] [Google Scholar]
  • 217.Brumme Z, Chan K, Dong W, et al. Prevalence and clinical implications of insertions in the HIV-1 p6Gag N-terminal region in drug-naive individuals initiating ART. Antivir Ther. 2003;8:91–6. [PubMed] [Google Scholar]
  • 218.Myint L, Matsuda M, Matsuda Z, et al. Gag non-cleavage site mutations contribute to full recovery of viral fitness in PI-resistant HIV-1. Antimicrob Agents Chemother. 2004;48:444–52. doi: 10.1128/AAC.48.2.444-452.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219.Lastere S, Dalban C, Collin G, et al. Impact of insertions in the HIV-1 p6 PTAPP region on the virologic response to amprenavir. Antivir Ther. 2004;9:221–7. [PubMed] [Google Scholar]
  • 220.Parkin N, Schapiro J. Antiretroviral drug resistance in non-subtype B HIV-1, HIV-2 and SIV. Antivir Ther. 2004;9:3–12. [PubMed] [Google Scholar]
  • 221.Rhee S, Kantor R, Katzenstein D, et al. HIV-1 pol mutation frequency by subtype and treatment experience: extension of the HIVseq program to seven non-B subtypes. AIDS. 2006;20:643–51. doi: 10.1097/01.aids.0000216363.36786.2b. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Velazquez-Campoy A, Todd M, Vega S, Freire E. Catalytic efficiency and vitality of HIV-1 proteases from African viral subtypes. Proc Natl Acad Sci USA. 2001;98:6062–7. doi: 10.1073/pnas.111152698. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Holguin A, Paxinos E, Hertogs K, Womac C, Soriano V. Impact of frequent natural polymorphisms at the protease gene on the in vitro susceptibility to PI in HIV-1 non-B subtypes. J Clin Virol. 2004;31:215–20. doi: 10.1016/j.jcv.2004.03.015. [DOI] [PubMed] [Google Scholar]
  • 224.Sanches M, Krauchenco S, Martins N, Gustchina A, Wlodawer A, Polikarpov I. Structural characterization of B and non-B subtypes of HIV-protease: insights into the natural susceptibility to drug resistance development. J Mol Biol. 2007;369:1029–40. doi: 10.1016/j.jmb.2007.03.049. [DOI] [PubMed] [Google Scholar]
  • 225.Geretti A. HIV-1 subtypes: epidemiology and significance for HIV management. Curr Opin Infect Dis. 2006;19:1–7. doi: 10.1097/01.qco.0000200293.45532.68. [DOI] [PubMed] [Google Scholar]
  • 226.Kantor R, Katzenstein D, Efron B, 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. doi: 10.1371/journal.pmed.0020112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Cane P, de Ruiter A, Rice P, Wiselka M, Fox R, Pillay D. Resistance-associated mutations in the HIV-1 subtype c protease gene from treated and untreated patients in the UK. J Clin Microbiol. 2001;39:2652–4. doi: 10.1128/JCM.39.7.2652-2654.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 228.Sugiura W, Matsuda Z, Yokomaku Y, et al. Interference between D30N and L90M in selection and development of PI-resistant HIV-1. Antimicrob Agents Chemother. 2002;46:708–15. doi: 10.1128/AAC.46.3.708-715.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229.Grossman Z, Paxinos E, Averbuch D, et al. Mutation D30N is not preferentially selected by HIV-1 subtype C in the development of resistance to nelfinavir. Antimicrob Agents Chemother. 2004;48:2159–65. doi: 10.1128/AAC.48.6.2159-2165.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 230.Abecasis A, Deforche K, Snoeck J, et al. Protease mutation M89I/V is linked to therapy failure in patients infected with the HIV-1 non-B subtypes C, F or G. AIDS. 2005;19:1799–806. doi: 10.1097/01.aids.0000188422.95162.b7. [DOI] [PubMed] [Google Scholar]
  • 231.Calazans A, Brindeiro R, Brindeiro P, et al. Low accumulation of L90M in protease from subtype F HIV-1 with resistance to PI is caused by the L89M polymorphism. J Infect Dis. 2005;191:1961–70. doi: 10.1086/430002. [DOI] [PubMed] [Google Scholar]
  • 232.Gonzalez L, Brindeiro R, Aguiar R, et al. Antimicrob Agents Chemother. Vol. 48. 2004. Impact of nelfinavir resistance mutations on in vitro phenotype, fitness, and replication capacity of HIV-1 with subtype B and C proteases; pp. 3552–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 233.Deforche K, Silander T, Camacho R, et al. Analysis of HIV-1 pol sequences using Bayesian networks: implications for drug resistance. Bioinformatics. 2006;22:2975–9. doi: 10.1093/bioinformatics/btl508. [DOI] [PubMed] [Google Scholar]
  • 234.Camacho R, Godinho A, Gomes P, et al. Different substitutions under drug pressure at protease codon 82 in HIV-1 subtype G compared to subtype B infected individuals including a novel I82M resistance mutations. Antivir Ther. 2005;10:S151. abstract 138. [Google Scholar]
  • 235.van de Vijver D, Wensing A, Angarano G, et al. The calculated genetic barrier for antiretroviral drug resistance substitutions is largely similar for different HIV-1 subtypes. J Acquir Immune Defic Syndr. 2006;41:352–60. doi: 10.1097/01.qai.0000209899.05126.e4. [DOI] [PubMed] [Google Scholar]
  • 236.van Gent D, Groeneger A, Plasterk R. Mutational analysis of the integrase protein of HIV- 2. Proc Natl Acad Sci USA. 1992;89:9598–602. doi: 10.1073/pnas.89.20.9598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 237.Pommier Y, Johnson A, Marchand C. Integrase inhibitors to treat HIV/AIDS. Nat Rev Drug Discov. 2005;4:236–48. doi: 10.1038/nrd1660. [DOI] [PubMed] [Google Scholar]
  • 238.Chen J, Krucinski J, Miercke L, et al. Crystal structure of the HIV-1 integrase catalytic core and C-terminal domains: a model for viral DNA binding. Proc Natl Acad Sci USA. 2000;97:8233–8. doi: 10.1073/pnas.150220297. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 239.Wang J, Ling H, Yang W, Craigie R. Structure of a two-domain fragment of HIV-1 integrase: implications for domain organization in the intact protein. Embo J. 2001;20:7333–43. doi: 10.1093/emboj/20.24.7333. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 240.Goldgur Y, Craigie R, Cohen G, et al. Structure of the HIV-1 integrase catalytic domain complexed with an inhibitor: a platform for antiviral drug design. Proc Natl Acad Sci USA. 1999;96:13040–3. doi: 10.1073/pnas.96.23.13040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 241.Fikkert V, Hombrouck A, Van Remoortel B, et al. Multiple mutations in HIV-1 integrase confer resistance to the clinical trial drug S-1360. AIDS. 2004;18:2019–28. doi: 10.1097/00002030-200410210-00006. [DOI] [PubMed] [Google Scholar]
  • 242.Hazuda D, Anthony N, Gomez R, et al. A naphthyridine carboxamide provides evidence for discordant resistance between mechanistically identical inhibitors of HIV-1 integrase. Proc Natl Acad Sci USA. 2004;101:11233–8. doi: 10.1073/pnas.0402357101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 243.Sotriffer C, Ni H, McCammon J. Active site binding modes of HIV-1 integrase inhibitors. J Med Chem. 2000;43:4109–17. doi: 10.1021/jm000194t. [DOI] [PubMed] [Google Scholar]
  • 244.Savarino A. In-silico docking of HIV-1 integrase inhibitors reveals a novel drug type acting on an enzyme/DNA reaction intermediate. Retrovirology. 2007;4:21. doi: 10.1186/1742-4690-4-21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 245.Shimura K, Kodama E, Sakagami Y, et al. Broad antiretroviral activity and resistance profile of a novel HIV integrase inhibitor, elvitegravir (JTK303/GS-9137) J Virol. 2008;82:764–74. doi: 10.1128/JVI.01534-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 246.Suzuki H, Fujino M, Matsuda M, Yen H, Iwatani Y, Sugiura W. Effects of protease and reverse transcriptase inhibitor-resistance mutations on integrase polymorphisms in multidrug resistance cases. Antiviral Ther. 2007;12:S4. abstract 2. [Google Scholar]
  • 247.Myers R, Pillay D. HIV-1 integrase sequence variation and covariation. Antiviral Ther. 2007;12:S5. abstract 3. [Google Scholar]
  • 248.Ceccherini-Silberstein F, Malet I, Fabeni L, et al. Specific mutations related to resistance to HIV-1 integrase inhibitors are associated with reverse transcriptase mutations in HAART-treated patients. Antivir Ther. 2007;12:S6. abstract 4. [Google Scholar]
  • 249.Lataillade M, Chiarella J, Kozal M. Natural polymorphism of the HIV-1 integrase gene and mutations associated with integrase inhibitor resistance. Antivir Ther. 2007;12:563–70. [PubMed] [Google Scholar]
  • 250.Jones G, Ledford R, Yu F, Miller M, Tsiang M, McColl D. Resistance profile of HIV-1 mutants in vitro selected by the HIV-1 integrase inhibitor, GS-9137 (JTK-303). 14th CROI; Los Angeles. 2007. abstract 627. [Google Scholar]
  • 251.Hazuda D, Miller M, Nguyen B, Zhao J. Resistance to the HIV integrase inhibitor raltegravir: analysis of protocol 005, a phase II study in patients with triple-class resistant HIV-1. Antivir Ther. 2007;12 abstract 8. [Google Scholar]
  • 252.Mccoll D, Fransen S, Gupta S, et al. Resistance and cross-resistance to first generation integrase inhibitors: insights from a phase II study of elvitegravir (GS-9137) Antivir Ther. 2007;12 abstract 9. [Google Scholar]
  • 253.Ren C, May S, Miletti T, Bedard J. In vitro cross-resistance studies of five different classes of integrase inhibitors in recombinant HIV-1. Antivir Ther. 2007;12:S3. abstract 1. [Google Scholar]
  • 254.Hombrouck A, Voet A, Van Remoortel B, et al. Mutations in HIV-1 integrase confer resistance to the naphthyridine L-870,810 and cross-resistance to the clinical trial drug GS-9137. Antimicrob Agents Chemother. 2008;52:2069–78. doi: 10.1128/AAC.00911-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 255.Rowley M. The discovery of raltegravir, an integrase inhibitor for the treatment of HIV infection. Prog Med Chem. 2008;46:1–28. doi: 10.1016/S0079-6468(07)00001-X. [DOI] [PubMed] [Google Scholar]
  • 256.Merck. Isentress. Package Insert 2007.
  • 257.Malet I, Delelis O, Valantin M, et al. Mutations associated with failure of raltegravir treatment affect integrase sensitivity to the inhibitor in vitro. Antimicrob Agents Chemother. 2008;52:1351–8. doi: 10.1128/AAC.01228-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 258.Cabrera C, Marfil S, Garcia E, et al. Genetic evolution of gp41 reveals a highly exclusive relationship between codons 36, 38 and 43 in gp41 under long-term enfuvirtide-containing salvage regimen. AIDS. 2006;20:2075–80. doi: 10.1097/QAD.0b013e3280102377. [DOI] [PubMed] [Google Scholar]
  • 259.Lu J, Deeks S, Hoh R, et al. Rapid emergence of enfuvirtide resistance in HIV-1-infected patients: results of a clonal analysis. J Acquir Immune Defic Syndr. 2006;43:60–4. doi: 10.1097/01.qai.0000234083.34161.55. [DOI] [PubMed] [Google Scholar]
  • 260.Hanna S, Yang C, Owen S, Lal R. Variability of critical epitopes within HIV-1 heptad repeat domains for selected entry inhibitors in HIV-infected populations worldwide [corrected] AIDS. 2002;16:1603–8. doi: 10.1097/00002030-200208160-00005. [DOI] [PubMed] [Google Scholar]
  • 261.Xu L, Hue S, Taylor S, et al. Minimal variation in T-20 binding domain of different HIV-1 subtypes from antiretroviral-naive and -experienced patients. AIDS. 2002;16:1684–6. doi: 10.1097/00002030-200208160-00016. [DOI] [PubMed] [Google Scholar]
  • 262.Villahermosa M, Perez-Alvarez L, Carmona R, et al. Primary resistance mutations to fusion inhibitors and polymorphisms in gp41 sequences of HIV-1 non-B subtypes and recombinants. AIDS. 2003;17:1083–6. doi: 10.1097/00002030-200305020-00020. [DOI] [PubMed] [Google Scholar]
  • 263.Cilliers T, Patience T, Pillay C, Papathanasopoulos M, Morris L. Sensitivity of HIV type 1 subtype C isolates to the entry inhibitor T-20. AIDS Res Hum Retroviruses. 2004;20:477–82. doi: 10.1089/088922204323087714. [DOI] [PubMed] [Google Scholar]
  • 264.Aghokeng A, Ewane L, Awazi B, et al. Enfuvirtide binding domain is highly conserved in non-B HIV-1 strains from Cameroon, West Central Africa. AIDS Res Hum Retroviruses. 2005;21:430–3. doi: 10.1089/aid.2005.21.430. [DOI] [PubMed] [Google Scholar]
  • 265.Holguin A, Faudon J, Labernardiere J, Soriano V. Susceptibility of HIV-1 non-B subtypes and recombinant variants to enfuvirtide. J Clin Virol. 2007;38:176–80. doi: 10.1016/j.jcv.2006.09.002. [DOI] [PubMed] [Google Scholar]
  • 266.Labrosse B, Labernardiere J, Dam E, et al. Baseline susceptibility of primary HIV-1 to entry inhibitors. J Virol. 2003;77:1610–3. doi: 10.1128/JVI.77.2.1610-1613.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 267.Sista P, Melby T, Greenberg M, et al. Subgroup analysis of baseline susceptibility and early virologic response to enfuvirtide in the combined TORO studies. Antivir Ther. 2003;8:S60. [Google Scholar]
  • 268.Reeves J, Gallo S, Ahmad N, et al. Sensitivity of HIV-1 to entry inhibitors correlates with envelope/coreceptor affinity, receptor density, and fusion kinetics. Proc Natl Acad Sci USA. 2002;99:16249–54. doi: 10.1073/pnas.252469399. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 269.Melby T, Sista P, DeMasi R, et al. Characterization of envelope glycoprotein gp41 genotype and phenotypic susceptibility to enfuvirtide at baseline and on treatment in the phase III clinical trials TORO-1 and TORO-2. AIDS Res Hum Retroviruses. 2006;22:375–85. doi: 10.1089/aid.2006.22.375. [DOI] [PubMed] [Google Scholar]
  • 270.Sista P, Melby T, Davison D, et al. Characterization of determinants of genotypic and phenotypic resistance to enfuvirtide in baseline and on-treatment HIV-1 isolates. AIDS. 2004;18:1787–94. doi: 10.1097/00002030-200409030-00007. [DOI] [PubMed] [Google Scholar]
  • 271.Menzo S, Castagna A, Monachetti A, et al. Genotype and phenotype patterns of HIV-1 resistance to enfuvirtide during long-term treatment. Antimicrob Agents Chemother. 2004;48:3253–9. doi: 10.1128/AAC.48.9.3253-3259.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 272.Marcelin A, Reynes J, Yerly S, et al. Characterization of genotypic determinants in HR-1 and HR-2 gp41 domains in individuals with persistent HIV viremia under T-20. AIDS. 2004;18:1340–2. doi: 10.1097/00002030-200406180-00015. [DOI] [PubMed] [Google Scholar]
  • 273.Mink M, Mosier S, Janumpalli S, et al. Impact of HIV-1 gp41 amino acid substitutions selected during enfuvirtide treatment on gp41 binding and antiviral potency of enfuvirtide in vitro. J Virol. 2005;79:12447–54. doi: 10.1128/JVI.79.19.12447-12454.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 274.Su C, Melby T, DeMasi R, Ravindran P, Heilek-Snyder G. Genotypic changes in HIV-1 envelope glycoproteins on treatment with the fusion inhibitor enfuvirtide and their influence on changes in drug susceptibility in vitro. J Clin Virol. 2006;36:249–57. doi: 10.1016/j.jcv.2006.03.007. [DOI] [PubMed] [Google Scholar]
  • 275.Xu L, Pozniak A, Wildfire A, et al. Emergence and evolution of enfuvirtide resistance following long-term therapy involves heptad repeat 2 mutations within gp41. Antimicrob Agents Chemother. 2005;49:1113–9. doi: 10.1128/AAC.49.3.1113-1119.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 276.Baldwin C, Sanders R, Deng Y, et al. Emergence of a drug-dependent HIV-1 variant during therapy with the T20 fusion inhibitor. J Virol. 2004;78:12428–37. doi: 10.1128/JVI.78.22.12428-12437.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 277.Tolstrup M, Selzer-Plon J, Laursen A, et al. Full fusion competence rescue of the enfuvirtide resistant HIV-1 gp41 genotype (43D) by a prevalent polymorphism (137K) AIDS. 2007;21:519–21. doi: 10.1097/QAD.0b013e3280187558. [DOI] [PubMed] [Google Scholar]
  • 278.Cilliers T, Moore P, Coetzer M, Morris L. In vitro generation of HIV type 1 subtype C isolates resistant to enfuvirtide. AIDS Res Hum Retroviruses. 2005;21:776–83. doi: 10.1089/aid.2005.21.776. [DOI] [PubMed] [Google Scholar]
  • 279.D'Arrigo R, Ciccozzi M, Gori C, et al. gp41 sequence variability in HIV type 1 non-B subtypes infected patients undergoing enfuvirtide pressure. AIDS Res Hum Retroviruses. 2007;23:1296–302. doi: 10.1089/aid.2007.0095. [DOI] [PubMed] [Google Scholar]
  • 280.Lu J, Sista P, Giguel F, Greenberg M, Kuritzkes D. Relative replicative fitness of HIV-1 mutants resistant to enfuvirtide (T-20) J Virol. 2004;78:4628–37. doi: 10.1128/JVI.78.9.4628-4637.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 281.Deeks S, Lu J, Hoh R, et al. Interruption of enfuvirtide in HIV-1 infected adults with incomplete viral suppression on an enfuvirtide-based regimen. J Infect Dis. 2007;195:387–91. doi: 10.1086/510531. [DOI] [PubMed] [Google Scholar]
  • 282.Aquaro S, D'Arrigo R, Svicher V, et al. Specific mutations in HIV-1 gp41 are associated with immunologic success in HIV-1-infected patients receiving enfuvirtide treatment. J Antimicrob Chemother. 2006;58:714–22. doi: 10.1093/jac/dkl306. [DOI] [PubMed] [Google Scholar]
  • 283.Reeves J, Lee F, Miamidian J, Jabara C, Juntilla M, Doms R. Enfuvirtide resistance mutations: impact on HIV envelope function, entry inhibitor sensitivity, and virus neutralization. J Virol. 2005;79:4991–9. doi: 10.1128/JVI.79.8.4991-4999.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 284.Hartley O, Klasse P, Sattentau Q, Moore J. V3: HIV's switch hitter. AIDS Res Hum Retroviruses. 2005;21:171–89. doi: 10.1089/aid.2005.21.171. [DOI] [PubMed] [Google Scholar]
  • 285.Dragic T, Trkola A, Thompson D, et al. A binding pocket for a small molecule inhibitor of HIV-1 entry within the transmembrane helices of CCR5. Proc Natl Acad Sci USA. 2000;97:5639–44. doi: 10.1073/pnas.090576697. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 286.Castonguay L, Weng Y, Adolfsen W, et al. Binding of 2-aryl-4-(piperidin-1yl)butanamines and 1,3,4-trisubstituted pyrrolidines to human CCR5: a molecular modeling-guided mutagenesis study of the binding pocket. Biochemistry. 2003;42:1544–50. doi: 10.1021/bi026639s. [DOI] [PubMed] [Google Scholar]
  • 287.Tsamis F, Gavrilov S, Kajumo F, et al. Analysis of the mechanism by which the small molecule CCR5 antagonists SCH-351125 and SCH-350581 inhibit HIV-1 entry. J Virol. 2003;77:5201–8. doi: 10.1128/JVI.77.9.5201-5208.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 288.Westby M, van der Ryst E. CCR5 antagonists: host-targeted antivirals for the treatment of HIV infection. Antivir Chem Chemother. 2005;16:339–54. doi: 10.1177/095632020501600601. [DOI] [PubMed] [Google Scholar]
  • 289.Maeda K, Das D, Ogata-Aoki H, et al. Structural and molecular interactions of CCR5 inhibitors with CCR5. J Biol Chem. 2006;281:12688–98. doi: 10.1074/jbc.M512688200. [DOI] [PubMed] [Google Scholar]
  • 290.Seibert C, Ying W, Gavrilov S, et al. Interaction of small molecule inhibitors of HIV-1 entry with CCR5. Virology. 2006;349:41–54. doi: 10.1016/j.virol.2006.01.018. [DOI] [PubMed] [Google Scholar]
  • 291.Lobritz M, Marozsan A, Troyer R, Arts E. Natural variation in the V3 crown of HIV-1 affects replicative fitness and entry inhibitor sensitivity. J Virol. 2007;81:8258–69. doi: 10.1128/JVI.02739-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 292.Pugach P, Marozsan A, Ketas T, Landes E, Moore J, Kuhmann S. HIV-1 clones resistant to a small molecule CCR5 inhibitor use the inhibitor-bound form of CCR5 for entry. Virology. 2007;361:212–28. doi: 10.1016/j.virol.2006.11.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 293.Dorr P, Westby M, Dobbs S, et al. Maraviroc (UK-427,857), a potent, orally bioavailable, and selective small-molecule inhibitor of chemokine receptor CCR5 with broad-spectrum anti-HIV-1 activity. Antimicrob Agents Chemother. 2005;49:4721–32. doi: 10.1128/AAC.49.11.4721-4732.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 294.Strizki J, Tremblay C, Xu S, et al. Discovery and characterization of vicriviroc (SCH 417690), a CCR5 antagonist with potent activity against HIV-1. Antimicrob Agents Chemother. 2005;49:4911–19. doi: 10.1128/AAC.49.12.4911-4919.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 295.Trkola A, Kuhmann S, Strizki J, et al. HIV-1 escape from a small molecule, CCR5-specific entry inhibitor does not involve CXCR4 use. Proc Natl Acad Sci USA. 2002;99:395–400. doi: 10.1073/pnas.012519099. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 296.Maeda K, Nakata H, Koh Y, et al. Spirodiketopiperazine-based CCR5 inhibitor which preserves CC-chemokine/CCR5 interactions and exerts potent activity against R5 HIV-1 in vitro. J Virol. 2004;78:8654–62. doi: 10.1128/JVI.78.16.8654-8662.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 297.Marozsan A, Kuhmann S, Morgan T, et al. Generation and properties of a HIV-1 isolate resistant to the small molecule CCR5 inhibitor, SCH417690 (SCH-D) Virology. 2005;338:182–99. doi: 10.1016/j.virol.2005.04.035. [DOI] [PubMed] [Google Scholar]
  • 298.Westby M, Smith-Burchnell C, Mori J, et al. Reduced maximal inhibition in phenotypic susceptibility assays indicates that viral strains resistant to the CCR5 antagonist maraviroc utilize inhibitor-bound receptor for entry. J Virol. 2007;81:2359–71. doi: 10.1128/JVI.02006-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 299.Kuhmann S, Pugach P, Kunstman KJ, et al. Genetic and phenotypic analyses of HIV-1 escape from a small-molecule CCR5 inhibitor. J Virol. 2004;78:2790–807. doi: 10.1128/JVI.78.6.2790-2807.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 300.Rusert P, Kuster H, Joos B, et al. Virus isolates during acute and chronic HIV-1 infection show distinct patterns of sensitivity to entry inhibitors. J Virol. 2005;79:8454–69. doi: 10.1128/JVI.79.13.8454-8469.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 301.Baba M, Miyake H, Wang X, Okamoto M, Takashima K. Isolation and characterization of HIV-1 resistant to the small-molecule CCR5 antagonist TAK 652. Antimicrob Agents Chemother. 2007;51:707–15. doi: 10.1128/AAC.01079-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 302.Ogert R, Wojcik L, Buontempo C, et al. Mapping resistance to the CCR5 co-receptor antagonist vicriviroc using heterologous chimeric HIV-1 envelope genes reveals key determinants in the C2-V5 domain of gp120. Virology. 2008;373:387–99. doi: 10.1016/j.virol.2007.12.009. [DOI] [PubMed] [Google Scholar]
  • 303.Mori J, Mosley M, Lewis M, et al. Characterization of maraviroc resistance in patients failing treatment with CCR5-tropic virus in MOTIVATE 1 and 2. Antivir Ther. 2007;12:S12. abstract 10. [Google Scholar]
  • 304.Tsibris A, Gulick R, Su Z, et al. In vivo emergence of HIV-1 resistance to the CCR5 antagonist vicriviroc: findings from ACTG 5211. Antivir Ther. 2007;12:S15. abstract 13. [Google Scholar]
  • 305.Moore J, Kitchen S, Pugach P, Zack J. The CCR5 and CXCR4 coreceptors-central to understanding the transmission and pathogenesis of HIV-1 infection. AIDS Res Hum Retroviruses. 2004;20:111–26. doi: 10.1089/088922204322749567. [DOI] [PubMed] [Google Scholar]
  • 306.Brumme Z, Goodrich J, Mayer H, et al. Molecular and clinical epidemiology of CXCR4 using HIV-1 in a large population of antiretroviral-naive individuals. J Infect Dis. 2005;192:466–74. doi: 10.1086/431519. [DOI] [PubMed] [Google Scholar]
  • 307.Hunt P, Harrigan P, Huang W, et al. Prevalence of CXCR4 tropism among antiretroviral-treated HIV-1-infected patients with detectable viremia. J Infect Dis. 2006;194:926–30. doi: 10.1086/507312. [DOI] [PubMed] [Google Scholar]
  • 308.Poveda E, Briz V, Quinones-Mateu M, Soriano V. HIV tropism: diagnostic tools and implications for disease progression and treatment with entry inhibitors. AIDS. 2006;20:1359–67. doi: 10.1097/01.aids.0000233569.74769.69. [DOI] [PubMed] [Google Scholar]
  • 309.Melby T, Despirito M, Demasi R, Heilek-Snyder G, Greenberg M, Graham N. HIV-1 coreceptor use in triple-class treatment-experienced patients: baseline prevalence, correlates, and relationship to enfuvirtide response. J Infect Dis. 2006;194:238–46. doi: 10.1086/504693. [DOI] [PubMed] [Google Scholar]
  • 310.Wilkin T, Su Z, Kuritzkes D, et al. HIV-1 chemokine coreceptor use among antiretroviral-experienced patients screened for a clinical trial of a CCR5 inhibitor: AIDS Clinical Trial Group A5211. Clin Infect Dis. 2007;44:591–5. doi: 10.1086/511035. [DOI] [PubMed] [Google Scholar]
  • 311.Whitcomb J, Huang W, Fransen S, et al. Development and characterization of a novel single-cycle recombinant-virus assay to determine HIV-1 coreceptor tropism. Antimicrob Agents Chemother. 2007;51:566–75. doi: 10.1128/AAC.00853-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 312.Huang W, Eshleman S, Toma J, et al. Coreceptor tropism in HIV-1 subtype D: high prevalence of CXCR4 tropism and heterogeneous composition of viral populations. J Virol. 2007;81:7885–93. doi: 10.1128/JVI.00218-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 313.Pastore C, Nedellec R, Ramos A, et al. Conserved changes in envelope function during HIV-1 coreceptor switching. J Virol. 2007;81:8165–79. doi: 10.1128/JVI.02792-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 314.Pastore C, Nedellec R, Ramos A, Pontow S, Ratner L, Mosier D. HIV-1 coreceptor switching: V1/V2 gain-of-fitness mutations compensate for V3 loss-of-fitness mutations. J Virol. 2006;80:750–8. doi: 10.1128/JVI.80.2.750-758.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 315.Pastore C, Ramos A, Mosier D. Intrinsic obstacles to HIV-1 coreceptor switching. J Virol. 2004;78:7565–74. doi: 10.1128/JVI.78.14.7565-7574.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 316.Jensen M, van 't Wout A. Predicting HIV-1 coreceptor usage with sequence analysis. AIDS Rev. 2003;5:104–12. [PubMed] [Google Scholar]
  • 317.Sander O, Sing T, Sommer I, et al. Structural descriptors of gp120 V3 loop for the prediction of HIV-1 coreceptor usage. PLoS Comput Biol. 2007;3:e58. doi: 10.1371/journal.pcbi.0030058. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 318.Sing T, Low A, Beerenwinkel N, et al. Predicting HIV coreceptor usage on the basis of genetic and clinical covariates. Antivir Ther. 2007;12:1097–106. [PubMed] [Google Scholar]
  • 319.Huang W, Toma J, Fransen S, et al. Coreceptor tropism can be influenced by amino acid substitutions in the gp41 transmembrane subunit of HIV-1 env protein. J Virol. 2008;82:5584–93. doi: 10.1128/JVI.02676-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 320.Moncunill G, Armand-Ugon M, Pauls E, Clotet B, Este J. HIV-1 escape to CCR5 coreceptor antagonism through selection of CXCR4-using variants in vitro. AIDS. 2008;22:23–31. doi: 10.1097/QAD.0b013e3282f303e6. [DOI] [PubMed] [Google Scholar]
  • 321.Reeves J, Han D, Wilkin T, et al. An enhanced version of the Trofile HIV coreceptor tropism assay predicts emergence of CXCR4 use in ACTG5211 vicriviroc trial samples; 15th CROI; 2008. abstract 869. [Google Scholar]
  • 322.Fatkenheuer G, Pozniak A, Johnson M, et al. Efficacy of short-term monotherapy with maraviroc, a new CCR5 antagonist, in patients infected with HIV-1. Nat Med. 2005;11:1170–2. doi: 10.1038/nm1319. [DOI] [PubMed] [Google Scholar]
  • 323.Westby M, Lewis M, Whitcomb J, et al. Emergence of CXCR4-using HIV-1 variants in a minority of HIV-1-infected patients following treatment with the CCR5 antagonist maraviroc is from a pretreatment CXCR4-using virus reservoir. J Virol. 2006;80:4909–20. doi: 10.1128/JVI.80.10.4909-4920.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 324.Lewis M, Simpson P, Fransen S, et al. CXCR4-using virus detected in patients receiving maraviroc in the phase III studies MOTIVATE 1 and 2 originates from a pre-existing minority of CXCR4-using virus. Antviral Ther. 2007;12:S65. abstract 56. [Google Scholar]
  • 325.Schurmann D, Fatkenheuer G, Reynes J, et al. Antiviral activity, pharmacokinetics and safety of vicriviroc, an oral CCR5 antagonist, during 14-day monotherapy in HIV-infected adults. AIDS. 2007;21:1293–9. doi: 10.1097/QAD.0b013e3280f00f9f. [DOI] [PubMed] [Google Scholar]
  • 326.Kitrinos K, Irlbeck D, LaBranche C, Madsen H, Demarest J. Virologic characterization of treatment naive subjects failing an aplaviroc-based regimen with either lamivudine/zidovudine or lopinavir/ritonavir. Antivir Ther. 2006;11:S26. abstract 21. [Google Scholar]
  • 327.Low A, Dong W, Chan D, et al. Current V3 genotyping algorithms are inadequate for predicting X4 co-receptor usage in clinical isolates. AIDS. 2007;21:F17–24. doi: 10.1097/QAD.0b013e3282ef81ea. [DOI] [PubMed] [Google Scholar]
  • 328.Wang C, Mitsuya Y, Gharizadeh B, Ronaghi M, Shafer RW. Characterization of mutation spectra with ultra-deep pyrosequencing: application to HIV-1 drug resistance. Genome Res. 2007;17:1195–201. doi: 10.1101/gr.6468307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 329.Tsibris A, Arnaout R, Lo C, et al. V3 loop sequence dynamics in subjects failing CCR5 antagonist therapy. Keystone Symposium on HIV Pathogenesis; Banff, Alberta, Canada. 2008. [Google Scholar]

RESOURCES