Newer
Older
grobid-corpus / fulltext / istex / tei / nature-1E01833F8B36A0ACBE48AC46A9B1A19EC2942043.training.fulltext.tei.xml
@zeynalig zeynalig on 26 Apr 2017 51 KB initialisation des corpus
<?xml version="1.0" ?>
<tei>
	<teiHeader>
		<fileDesc xml:id="0"/>
	</teiHeader>
	<text xml:lang="en">

		<p>Mammalian cells encounter a variety of DNA-damaging agents and<lb/> possess a number of
			different response pathways to maintain genomic<lb/> integrity <ref type="biblio"
				>1</ref> . One of the less well defined DNA damage responses results<lb/> from
			retroviral infections such as those induced by HIV-1. Integration<lb/> of the
			double-stranded cDNA reverse transcriptase product into the<lb/> host genome is an
			essential step in the retroviral life cycle. This event is<lb/> catalysed in part by the
			viral integrase protein that cleaves the host DNA<lb/> and promotes a strand transfer
			reaction resulting in short staggered DNA<lb/> breaks at the site of attachment <ref
				type="biblio">2</ref> . The gapped DNA intermediates then<lb/> have to be detected
			and efficiently repaired by host cell proteins if there<lb/> is to be a completed
			integration process and a productive infection. In<lb/> addition, it has been suggested
			that the unintegrated linear viral cDNA<lb/> may itself invoke a DNA damage response
			that, if left unchecked, will<lb/> result in the death of the host cell. This raises the
			possibility that the<lb/> inhibition of host DNA damage response factors could provide a
			new<lb/> therapeutic approach for the treatment of HIV-1 infections, especially<lb/>
			because some of these factors are not essential for cell survival.<lb/></p>

		<p>A number of groups have provided evidence that the Ku-dependent<lb/> non-homologous
			end-joining (NHEJ) pathway that is normally associ-<lb/>ated with the repair of DNA
			double-strand breaks (DSBs) is required<lb/> to support efficient retroviral infection
				<ref type="biblio">3–6</ref> and to prevent viral-induced<lb/> cell death <ref
				type="biblio">3,4</ref> . However, there is some controversy as to whether cell<lb/>
			death following retroviral infection is dependent on integrase activity <ref
				type="biblio">3,4</ref> .<lb/> Additional questions have also been raised about the
			involvement of<lb/> the ATM and ataxia-telangiectasia-and Rad3-related (ATR)
			proteins<lb/> in retroviral-induced DNA damage responses <ref type="biblio">7,8</ref> .
			Both ATM and ATR<lb/> are phosphatidylinositol-3-OH-kinase (PI(3)K)-like
			serine/threonine<lb/> kinases (PIKKs) that regulate cellular responses to DNA damage by
			con-<lb/>trolling cell-cycle arrest and DNA repair pathways <ref type="biblio"
				>9,10</ref> . Although many of<lb/> the protein substrates of ATM and ATR overlap
				<ref type="biblio">9</ref> , activation of the ATM<lb/> or ATR kinases largely
			depends on the type of DNA damage lesion and<lb/> the stage of the cell cycle in which
			it is encountered. ATM predomi-<lb/>nantly detects DNA DSBs, such as those caused by
			ionizing radiation,<lb/> and responds very rapidly to these lesions at all stages of the
			cell cycle,<lb/> whereas ATR is activated by agents such as ultraviolet radiation
			and<lb/> replication-fork collapse during S phase <ref type="biblio">10</ref> .<lb/></p>

		<p>Here, by using both genetic and pharmacological approaches, we dem-<lb/>onstrate that ATM
			activity has an important role in retroviral replication.<lb/> By screening a small
			molecule compound library developed for the PIKK<lb/> family, we have identified
			KU-55933, a novel, specific and potent inhibi-<lb/>tor of ATM <ref type="biblio"
				>11</ref> . This compound, which is not effective in targeting ATR,<lb/>
			DNA-dependent protein kinase (DNA-PK) or PI(3)K, suppresses HIV-1<lb/> replication and
			provides important proof of concept that the inhibition<lb/> of ATM, a non-essential
			cellular target, may represent a new approach<lb/> to the treatment of HIV-1
			infections.<lb/></p>

		<head>RESULTS<lb/></head>

		<head> ATM facilitates efficient HIV-1 infection<lb/></head>

		<p>To investigate whether the absence of ATM affects HIV-1 infection,<lb/> both ATM knockout
			mouse embryonic stem (ES) cells and human<lb/> fibroblast cells from ataxia
			telangiectasia (AT) patients were tested for<lb/> their ability to support HIV-1
			integrase-proficient (HIV-1 IN + ) vector<lb/> transduction <ref type="figure">(Fig.
				1a)</ref> using both a colony formation assay based on<lb/> antibiotic selection and
			a luciferase reporter gene assay <ref type="biblio">12</ref> . For both the<lb/> ATM
			knockout mouse ES cells and the human AT fibroblast cells the<lb/> results demonstrated
			a clear reduction in efficiency of HIV-1 vector<lb/> transduction compared with the
			control cells expressing functional<lb/> ATM. To confirm these results, we performed
			identical assays using<lb/> previously described AT fibroblast cells <ref type="biblio"
				>13</ref> that had been complemented<lb/> with an ATM expression vector <ref
				type="figure">(Fig. 1b)</ref>. Consistent with the results<lb/> in <ref
				type="figure">Fig. 1a</ref>, ATM expression resulted in higher transduction rates
			com-<lb/>pared with the AT cells that contained vector alone. Because these<lb/>
			findings were different from those described in a previous study <ref type="biblio"
				>7</ref> , we<lb/> wished to extend these experiments from single-step
			transduction<lb/> assays to an investigation of wild-type HIV-1 replication.
			Replication-<lb/>competent wild-type HIV-1 virus was used to infect lymphoblast
			cells<lb/> derived from AT patients (GM01526) or control cells that contained<lb/>
			functional ATM (GM14680). HIV-1 replication was assessed by meas-<lb/>uring the levels
			of viral p24 antigen that were present in cell-free super-<lb/>natants after 7 days <ref
				type="figure">(Fig. 1c)</ref>. The reduced levels of viral replication<lb/> in
			ATM-deficient lymphoblasts, together with the results from the<lb/> transduction assays,
			support the idea that ATM has an important role<lb/> in HIV-1 infections. This role can
			probably be extended to retroviruses<lb/> in general because a similar dependency on ATM
			was seen for viral<lb/> vector transduction assays that used murine embryonic stem cell
			virus<lb/> (see <ref type="figure">Supplementary Information, Fig. S1</ref>).<lb/></p>

		<head>HIV-1 infection activates the ATM-dependent DNA damage<lb/> response
			pathway<lb/></head>

		<p>There are a number of studies that describe both the activation of ATM<lb/> in response
			to ionizing radiation-induced DSBs <ref type="biblio">9,10,14,15</ref> and the
			subsequent<lb/> phosphorylation of cellular targets. We investigated the activation of
			ATM<lb/></p>

		<figure>b<lb/> 0<lb/> 1<lb/> 2<lb/> 3<lb/> 4<lb/> AT<lb/> ATM<lb/> complemented<lb/> a<lb/>
			0<lb/> 20<lb/> 40<lb/> 60<lb/> 80<lb/> 100<lb/> 120<lb/> ATM +/+<lb/> 1BR<lb/> AT5<lb/>
			Human fibroblast<lb/> c<lb/> GM14680<lb/> (WT ATM)<lb/> GM01526<lb/> (AT)<lb/> 0<lb/>
			3.5<lb/> 3.0<lb/> 2.5<lb/> 2.0<lb/> 1.5<lb/> 1.0<lb/> 0.5<lb/> Transduction efficiency
			(%)<lb/> Mouse ES<lb/> ATM −/−<lb/> Relative transduction efficiency<lb/> HIV p24 (pg
			ml<lb/> −1 ) (×10<lb/> 4 )<lb/> Figure 1 ATM function is required for efficient HIV-1
			infection. (a) Retroviral<lb/> transduction assays using HIV-1 IN + viral vectors in
			7–10-day colony formation<lb/> (hatched), or 3-day luciferase assays (solid), were
			performed for matched<lb/> ATM −/− knockout mouse ES cells, and also with AT5 human AT
			fibroblasts<lb/> or 1BR fibroblasts containing wild-type ATM. Data are shown as the
			mean<lb/> percentage of viral transduction (transduction efficiency) compared with
			ATM-<lb/>expressing control cells (± s.d. from at least two independent experiments).
			For<lb/> all data in Fig. 1, wild-type or complemented control cells are shown in
			blue<lb/> and ATM-deficient cells are shown in orange. (b) Colony formation
			(hatched)<lb/> and luciferase (solid) HIV-1 IN + vector transduction assays were
			performed<lb/> using matched AT22IJE-T human AT fibroblast cells (AT22IJE-T pEBS7)
			and<lb/> AT cells complemented with the ATM cDNA (AT22IJE-T pEBS7-YZ5). Data<lb/> are
			shown as relative viral transduction efficiency compared with the AT cells<lb/> (± s.d.
			from two independent experiments). (c) HIV-1 replication assays for<lb/> GM14680
			wild-type (WT) and GM01526 AT human lymphoblast cells infected<lb/> with HIV-1 RF
			wild-type virus strain. Data are shown for the mean amount of<lb/> HIV-1 p24 antigen
			present in cell-free supernatants 7 days after infection (±<lb/> s.d. from three
			replicate infections.<lb/> a<lb/> b<lb/></figure>

		<figure>Time after addition (h)<lb/> Virus<lb/> IR<lb/> HIV-1 IN +<lb/> HIV-1 IN +<lb/>
			HIV-1 IN D64V<lb/> HIV-1 IN D64V<lb/> HIV-1 IN +<lb/> HIV-1 IN D64V<lb/> p53<lb/>
			β-actin<lb/> Ser 15 p53<lb/> P<lb/> β-actin<lb/> ATR<lb/> DNA-PK<lb/> ATM<lb/> Ser 1981
			ATM<lb/> P<lb/> p53<lb/> Ser 15 p53<lb/> P<lb/> NBS1<lb/> Ser 345 NBS1<lb/> P<lb/>
			GM14680<lb/> (WT ATM)<lb/> GM01526<lb/> (AT)<lb/> CHK2<lb/> Thr 68 CHK2<lb/> P<lb/> U
			T<lb/> H IV -1 IN +<lb/> H IV -1 IN D 6 4 V<lb/> IR<lb/> U V<lb/> U T<lb/> H IV -1 IN
			+<lb/> H IV -1 IN D 6 4 V<lb/> IR<lb/> U V<lb/> 0<lb/> 1<lb/> 2<lb/> 4<lb/> 6<lb/>
			8<lb/> 2 Gy<lb/> Figure 2 Functional HIV-1 integrase is required to elicit an
			ATM-dependent<lb/> DNA damage response. (a) DNA damage responses of GM14680
			wild-type<lb/> and GM01526 AT cells transduced with viral vectors. Results of
			whole-cell<lb/> lysate immunoblots from untreated cells (UT) or those infected with
			equal<lb/> amounts of integrase-proficient (HIV-1 IN + ) or D64V integrase mutant<lb/>
			(HIV-1 IN D64V ) vectors at 6 h post-infection. Also shown are immunoblots of<lb/> cells
			collected 1 h after exposure to 2 Gy ionizing radiation (IR) or 20 J m −2<lb/>
			ultraviolet radiation (UV). Immunoblots were sequentially probed with<lb/>
			nonphospho-and phospho-specific antibodies against ATM, CHK2, p53 and<lb/> NBS1.
			Immunoblots were also probed with nonphospho-specific antibodies<lb/> against ATR,
			DNA-PK and β-actin (loading controls). (b) Time course of p53<lb/> Ser 15
			phosphorylation after transduction with HIV-1 IN + and HIV-1 IN D64V<lb/> mutant viral
			vectors. Results of whole-cell lysate immunoblots from U2OS<lb/> cells infected with
			equal amounts of viral vectors at increasing time points<lb/> after infection. Also
			shown are immunoblots of U2OS lysates from cells<lb/> collected 15 min after exposure to
			2 Gy ionizing radiation. Immunoblots<lb/> were probed with antibodies specific to p53
			P-Ser 15 before re-probing for<lb/> total p53 or β-actin (loading
			controls).<lb/></figure>

		<p>after HIV-1 infection by monitoring a number of these well characterized<lb/>
			ATM-dependent phosphorylation events. We infected both GM14680<lb/> and GM01526 AT cells
			with either functional HIV-1 IN + luciferase vec-<lb/>tors or those containing the
			inactivating integrase mutation D64V (ref.<lb/> 16) (HIV-1 IN D64V ). We then monitored
			the status of ATM phosphoser-<lb/>ine 1981 (P-Ser 1981) <ref type="biblio">17</ref> and
			CHK2 phosphothreonine 68 (P-Thr 68) <ref type="biblio">18</ref> ,<lb/> both specific
			markers for ATM activation, as well as p53 P-Ser 15 <ref type="biblio">(refs 19,<lb/>
				20)</ref> and NBS1 P-Ser 343 <ref type="biblio">(ref. 21)</ref>, which can be
			phosphorylated by both<lb/> ATM or ATR depending on the type of DNA damage
			involved.<lb/></p>

		<p>To illustrate the presence or absence of functional ATM and ATR<lb/> activity in GM14680
			and GM01526 AT cells, both were exposed to either<lb/> ionizing radiation, which is
			predominantly signalled through ATM, or<lb/> ultraviolet radiation, which is signalled
			through ATR <ref type="biblio">10</ref> . GM14680 cells<lb/> showed phosphorylation of
			all tested substrates in response to either<lb/> ionizing radiation or ultraviolet
			radiation, whereas GM01526 AT cells<lb/> only demonstrated phosphorylation of p53 Ser 15
			and NBS1 Ser 343 in<lb/> response to ultraviolet radiation, confirming the loss of ATM
			function<lb/> but intact ATR activity in these cells <ref type="figure">(Fig.
			2a)</ref>.<lb/></p>

		<p>Analysis of cells at 6 h post-infection with HIV-1 IN + vectors showed<lb/> an increase
			in phosphorylation for all the DNA damage signalling tar-<lb/>gets tested in GM14680
			cells but not in GM01526 AT cells <ref type="figure">(Fig. 2a)</ref>,<lb/> indicating
			that DNA damage signalling in response to HIV-1 infec-<lb/>tion proceeds predominantly
			through activation of ATM kinase. In<lb/> addition, infection with HIV-1 IN D64V mutant
			vectors demonstrated<lb/> a reduction in phosphorylation for most targets when compared
			with<lb/> HIV-1 IN + vectors, suggesting that integrase-induced host DNA damage<lb/> but
			not unintegrated viral cDNA predominantly elicits this response.<lb/> Interestingly, an
			exception to this seems to be ATM auto-phosphoryla-<lb/>tion on Ser 1981, which is also
			enhanced during infection with HIV-1<lb/> IN D64V mutant vectors. The fact that this ATM
			phosphorylation event<lb/> is not translated into the phosphorylation of downstream ATM
			targets<lb/> such as CHK2, p53 or NBS1 is discussed later.<lb/></p>

		<p>To monitor activation of the ATM pathway by HIV-1 infection in a time<lb/> course study
			and in an additional cell type, we infected U2OS cells that have<lb/> functional ATM and
			p53 with either HIV-1 IN + or IN D64V mutant vectors and<lb/> monitored p53 P-Ser 15
			levels at various time points after infection <ref type="figure">(Fig. 2b)</ref>.<lb/>
			It can be seen that infection with the HIV-1 IN + vector induces the
			phospho-<lb/>rylation of p53 Ser 15, peaking in these cells at around 4–6 h
			post-infection, a<lb/> time consistent with integration and cleavage of the host DNA by
			integrase <ref type="biblio">2</ref> .<lb/> In contrast, no such response was observed
			for the integrase-defective HIV-1<lb/> vector, in agreement with the data for the
			GM14680 cells <ref type="figure">(Fig. 2a)</ref>.<lb/></p>

		<head>HIV-1 infection of ATM-deficient cells results in enhanced host<lb/> cell
			death<lb/></head>

		<p>To determine whether there was enhanced cell death associated with<lb/> HIV-1 infection
			in ATM-deficient cells, we monitored the matched<lb/> ATM +/+ and ATM −/− ES cells
			infected with IN + or IN D64V mutant HIV-1<lb/> viral vectors. Cell death was assessed
			by determining the percentage of<lb/> annexin-V-positive cells at time points up to 48 h
			post-infection. We<lb/> observed a marked increase in the level of host cell death when
			ATM-<lb/>deficient cells were infected with HIV-1 IN + vectors, but this was not<lb/>
			the case when integrase-deficient virus was used or when the host cells<lb/> possessed
			functional ATM <ref type="figure">(Fig. 3)</ref>. Together, the results presented in<lb/>
			<ref type="figure">Figs 1–3</ref> suggest that the ATM DNA damage response pathway
			facilitates<lb/> efficient retroviral infection by contributing to the survival of host
			cells<lb/> in response to integrase-induced DNA damage.<lb/></p>

		<head>Suppression of HIV-1 transduction by a small-molecule ATM inhibitor<lb/></head>

		<p>The genetic data described above suggests that ATM could represent a<lb/> valid target
			for the suppression of HIV-1 infection by small molecule<lb/> kinase inhibitors.
			KU-55933 (2-morpholin-4-yl-6-thianthren-1-yl-<lb/>pyran-4-one; <ref type="figure">Fig.
				4a</ref>) is a novel ATP competitive inhibitor of ATM with<lb/> an in vitro
			inhibitory concentration that gives half-maximal activity (IC 50 )<lb/> of 13 nM and
			cellular activity in the low micromolar range <ref type="biblio">11</ref> .
			Cellular<lb/> activity of KU-55933 has been demonstrated through both
			radiosensiti-<lb/>zation experiments and the abrogation of
			ionizing-radiation-dependent<lb/> phosphorylation of a range of known ATM targets
			including p53, γH2AX<lb/> and NBS1. Notably, this compound shows specificity with
			respect to the<lb/> inhibition of other PIKK family members. In particular, it is
			important<lb/> to note that the in vitro IC 50 of KU-55933 for ATR is greater than 100
			µM<lb/> and cellular ATR-dependent phosphorylation events are not inhibited by<lb/> the
			drug in response to ultraviolet-radiation-induced damage <ref type="biblio">11</ref>
			.<lb/></p>

		<p>To test the ability of KU-55933 to inhibit HIV-1 infection, we<lb/> exposed Jurkat
			lymphoblast cells to HIV-1 IN + luciferase vectors with<lb/> increasing concentrations
			of compound <ref type="figure">(Fig. 4b)</ref>. KU-55933 effectively<lb/> inhibited
			HIV-1 transduction in these experiments with an IC 50 value<lb/> of approximately 1 µM
				<ref type="figure">(Fig. 4b; left panel)</ref>. The significant reduction<lb/> in
			transduction efficiency could not be attributed to cytotoxic effects<lb/> of the
			compound, as demonstrated by the effect of drug alone on the<lb/> viability of the
			Jurkat cells <ref type="figure">(Fig. 4b; right panel)</ref>. We confirmed that<lb/></p>

		<figure>HIV-1 IN +<lb/> HIV-1 IN D64V<lb/> Inactivated<lb/> No virus<lb/> ATM +/+<lb/>
			0<lb/> 10<lb/> 20<lb/> 30<lb/> 40<lb/> Hours post-infection<lb/> ATM −/−<lb/> 0<lb/>
			10<lb/> 20<lb/> 30<lb/> 40<lb/> a<lb/> b<lb/> 0<lb/> 1 2<lb/> 2 4<lb/> 3 6<lb/> 4 8<lb/>
			Hours post-infection<lb/> 0<lb/> 1 2<lb/> 2 4<lb/> 3 6<lb/> 4 8<lb/> Annexin-V<lb/> +
			cells (%)<lb/> Annexin-V<lb/> + cells (%)<lb/> Figure 3 ATM-deficient cells show
			enhanced cell death when transduced<lb/> with recombinant HIV-1 IN + vectors. (a, b)
			Cytopathicity studies in matched<lb/> ATM +/+ (a) and ATM −/− (b) mouse ES cells
			infected with HIV-1 vectors. Cells<lb/> were mock-infected (no virus; pink circles) or
			infected with equal amounts of<lb/> wild-type integrase (HIV-1 IN + ; blue squares),
			D64V integrase mutant (HIV-1<lb/> IN D64V ; orange triangles) or heat-inactivated IN +
			(inactivated; green crosses)<lb/> HIV-1 viral stocks. Cells were collected at increasing
			time points after<lb/> infection and annexin-V-stained cells were detected by flow
			cytometry. Data<lb/> are presented as the mean percentage of annexin-V-positive cells
			within the<lb/> total cell population ± s.d. from two independent
			experiments.<lb/></figure>

		<p>KU-55933 was able to inhibit an ATM-dependent response to viral-<lb/>induced DNA damage
			by infecting lymphoblast cells (GM14680) with<lb/> HIV-1 IN + luciferase vectors in the
			presence of 10 µM of compound<lb/> and monitoring the phosphorylation of p53 Ser 15 <ref
				type="figure">(Fig. 4c)</ref>. KU-55933<lb/> prevented this virus-dependent
			phosphorylation event and because the<lb/> compound does not demonstrate cellular
			activity against ATR, these<lb/> data provide further evidence that the phosphorylation
			of p53 Ser 15<lb/> after HIV-1 infection results from the activity of ATM. The use of
			KU-<lb/>55933 on both GM14680 and GM01526 AT cells <ref type="figure">(Fig. 4d)</ref>
			shows that<lb/> the ATM inhibitor reduces transduction efficiency of the GM14680<lb/>
			cells to a level that is close to that of the ATM-deficient GM01526 cells,<lb/> while
			having little effect on transduction in AT cells; this provides<lb/> further evidence
			that the anti-retroviral activity of KU-55933 results<lb/> predominantly from the
			inhibition of ATM.<lb/></p>

		<head>KU-55933 targets the post-integration DNA repair phase of the<lb/> retroviral life
			cycle<lb/></head>

		<p> Having demonstrated that KU-55933 inhibits HIV-1 vector transduc-<lb/> tion, we wanted
			to confirm that these results were not owing to the<lb/> inhibition of viral proteins
			such as reverse transcriptase or integrase.<lb/> To achieve this, 293 cells were
			transduced with HIV-1 IN + viral vec-<lb/>tors in the absence or presence of 10 µM
			KU-55933 or 1 µM zidovu-<lb/>dine (also known as AZT), a well characterized inhibitor of
			reverse<lb/> transcriptase. Total DNA was extracted from cells at increasing times<lb/>
			post-infection and the formation of late reverse transcriptase cDNA<lb/> (late-RT),
			2-LTR circle cDNA, as well as integrated proviral DNA,<lb/> were assessed by PCR <ref
				type="figure">(Fig. 5)</ref>. Comparative analysis of the formation<lb/> of late-RT
				<ref type="figure">(Fig. 5a)</ref> and 2-LTR circle <ref type="figure">(Fig.
				5b)</ref> HIV-1 cDNA shows<lb/> little effect of the presence of 10 µM KU-55933 but
			is reduced when<lb/> cells are treated with zidovudine. Analysis of integrated
			proviral<lb/> HIV-1 DNA through Alu-viral LTR PCR <ref type="figure">(Fig. 5c)</ref>
			also shows that<lb/> KU-55933 has little effect on integration up to 24 h after
			infection,<lb/> suggesting no direct activity against HIV-1 integrase. However, at<lb/>
			48 h after infection in the presence of KU-55933, the amount of<lb/> integrated HIV-1
			DNA is markedly reduced and this is consistent<lb/> with the loss of the ATM-inhibited
			infected cells through cell death<lb/>
			<ref type="figure">(Fig. 3)</ref>. Using purified integrase and reverse transcriptase
			proteins<lb/> in standard in vitro assays, KU-55933 was also shown to have no<lb/>
			inhibitory activity at concentrations up to 50 µM (data not shown).<lb/> Together, these
			data show that KU-55933 does not inhibit HIV-1<lb/> integrase or reverse transcriptase
			activities and the effects on HIV-1<lb/> vector transduction are consistent with ATM
			inhibition, resulting in<lb/> a failure to efficiently repair integration-induced DNA
			damage, and<lb/> ultimately leading to the death of the host cell.<lb/></p>

		<head>The ATM inhibitor KU-55933 suppresses HIV-1 replication<lb/></head>

		<p>The results that demonstrate the ability of KU-55933 to inhibit single-<lb/>round HIV-1
			vector transduction encouraged us to assess whether the<lb/> ATM inhibitor would also
			suppress non-attenuated virus replication.<lb/> We therefore compared the ability of
			KU-55933 to inhibit HIV-1 rep-<lb/>lication in C8166 human T-lymphocyte cells with that
			of the existing<lb/> anti-retroviral nucleoside reverse transcriptase inhibitors
			abacavir<lb/> and lamivudine (also known as 3TC). It can be seen from <ref type="figure"
				>Fig. 6</ref> that<lb/> KU-55933 was indeed able to suppress HIV-1 replication with
			an IC 50<lb/> value of 2.3 µM compared with 0.9 µM for abacavir and 0.2 µM for<lb/>
			lamivudine. None of the compounds demonstrated overt toxicity in<lb/> the C8166 cells,
			having half-maximal viability (CC 50 ) values greater<lb/> than 30 µM (data not
			shown).<lb/></p>

		<figure>a<lb/> O<lb/> O<lb/> O<lb/> N<lb/> S<lb/> S<lb/> KU-55933<lb/> 0<lb/> 0.4<lb/>
			0.8<lb/> 1.2<lb/> 1.6<lb/> 2.0<lb/> 2.4<lb/> d<lb/> c<lb/> − + − +<lb/> − +<lb/> −
			+<lb/> −<lb/> KU-55933:<lb/> −<lb/> 2<lb/> 4<lb/> 6<lb/> 8<lb/> Time after exposure
			(h)<lb/> β-actin<lb/> HIV-1 IN +<lb/> Ser 15 p53<lb/> P<lb/> 0<lb/> 20<lb/> 40<lb/>
			60<lb/> 80<lb/> 100<lb/> 120<lb/> KU-55933 (µM)<lb/> KU-55933 (µM)<lb/> KU-55933
			(µM)<lb/> 0<lb/> 20<lb/> 40<lb/> 60<lb/> 80<lb/> 100<lb/> 120<lb/> b<lb/> Transduction
			efficiency (%)<lb/> Viability (%)<lb/> 0<lb/> 2<lb/> 4<lb/> 6<lb/> 8<lb/> 1 0<lb/>
			0<lb/> 2<lb/> 4<lb/> 6<lb/> 8<lb/> 10<lb/> 0<lb/> 2<lb/> 4<lb/> 6<lb/> 8<lb/> 1 0<lb/> 2
			Gy IR<lb/> HIV-1 IN +<lb/> 2 Gy IR<lb/> Transduction luciferase activity (c.p.s.)
			(×10<lb/> 5 )<lb/> Figure 4 Suppression of HIV-1 vector transduction by KU-55933. (a)
			The<lb/> chemical structure of KU-55933. (b) HIV-1 IN + luciferase vector
			transduction<lb/> assays for Jurkat cells treated with KU-55933. Cells were infected
			with<lb/> HIV-1 luciferase vector 1 h before treatment with KU-55933 (left graph).
			In<lb/> parallel, Jurkat cells were mock infected (no virus) before treatment with
			KU-<lb/>55933 and viable cells determined by MTS assay (right graph). Data show<lb/> the
			mean percentage of luciferase activity (transduction efficiency) or viability<lb/>
			compared with untreated controls (± s.d. from two independent experiments).<lb/> (c)
			Immunoblots of GM14680 lymphoblast whole-cell lysates after infection<lb/> with HIV-1 IN
			+ luciferase vectors in the presence or absence of 10 µM KU-<lb/>55933. Lysates from
			ionizing radiation (IR)-treated cells are also shown for<lb/> comparison. Immunoblots
			were probed with antibodies against p53 P-Ser 15<lb/> and β-actin. (d) HIV-1 IN +
			luciferase vector transduction assays for GM14680<lb/> normal (blue diamonds) and
			GM01526 AT (orange squares) cells treated<lb/> with KU-55933. Data show mean luciferase
			activity (measured in counts per<lb/> second, c.p.s.) ± s.d. from 10
			replicates.<lb/></figure>

		<p>We were also interested to see how well KU-55933 would work in<lb/> combination with
			approved anti-HIV therapies such as the nucleoside<lb/> reverse transcriptase inhibitors
			or protease inhibitors. Using HIV-1 rep-<lb/>lication (cytopathicity) assays and
			standard methods for determining<lb/> drug synergism or antagonism <ref type="biblio"
				>22</ref> we constructed an isobologram plot for<lb/> KU-55933 and lamivudine <ref
				type="figure">(Fig. 6b)</ref>. The isobologram plot obtained for<lb/> KU-55933 and
			lamivudine demonstrates a synergistic inhibitory effect<lb/> on HIV-1 replication when
			using these two drugs in combination. A<lb/> similar finding was made in combination
			studies of KU-55933 with<lb/> zidovudine, and at least an additive effect was observed
			for the protease<lb/> inhibitor saquinavir (data not shown).<lb/></p>

		<head>KU-55933 inhibits drug-resistant HIV-1 as efficiently as<lb/> wild-type
			virus<lb/></head>

		<p>One of the key advantages of using an inhibitor that targets a cellular<lb/> factor
			required for the HIV-1 life cycle, rather than a viral protein, is that<lb/> it should
			be just as effective against drug-resistant HIV-1 strains that will<lb/> emerge after
			conventional anti-retroviral therapies. To test this principle<lb/> directly, we
			investigated the activity of KU-55933 against a number of<lb/> recombinant HIV-1 viruses
			containing mutations in their reverse tran-<lb/>scriptase genes known to be associated
			with drug resistance. <ref type="table">Table 1<lb/></ref> provides the IC 50 values of
			KU-55933, abacavir and lamivudine against<lb/> wild-type and drug-resistant HIV-1
			strains that contain defined muta-<lb/>tions in the reverse transcriptase gene in 6-day
			replication assays. It can<lb/> be seen that whereas the IC 50 values of abacavir and
			lamivudine are much<lb/> higher for the drug-resistant virus compared with the wild-type
			virus,<lb/> KU-55933 values do not show any significant changes and demonstrate<lb/>
			comparable efficacy against all mutant HIV-1 strains tested. Moreover,<lb/> the shift in
			the ranking of compound effectiveness against drug-resistant<lb/> strains of HIV-1, in
			which KU-55933 is above both abacavir and lami-<lb/>vudine, illustrates the potential
			value of a cellular target, particularly in<lb/> patients for whom existing
			anti-retroviral therapy has failed.<lb/></p>

		<head>KU-55933 suppresses replication of clinical isolates of HIV-1<lb/> in primary blood
			mononuclear cells<lb/></head>

		<p>In addition to testing KU-55933 against recombinant HIV-1 in a T-cell line,<lb/> we
			investigated the ability of the ATM inhibitor to suppress the replication<lb/> of
			clinical isolates of HIV-1 in primary blood mononuclear cells (PBMCs).<lb/> Data were
			obtained from 7-day replication assays based on viral p24 antigen<lb/> production. The
			IC 50 values for KU-55933 against the three wild-type clinical<lb/> HIV-1 isolates HIV-1
			CC15 , HIV-1 CC1 and HIV-1 105A are 1.7, 6.5 and 2.9 µM,<lb/> respectively. These values
			are comparable to previous results obtained with<lb/> recombinant HIV-1. The IC 50
			values for KU-55933 against the zidovudine-<lb/>and lamivudine-resistant clinical
			isolates of HIV-1 are 4.1 and 5.5 µM,<lb/> respectively, illustrating that the ATM
			inhibitor is just as effective against<lb/> these drug-resistant isolates of HIV-1 as it
			is against wild-type isolates.<lb/></p>

		<head>DISCUSSION<lb/></head>

		<p>The data obtained in this study demonstrate that ATM is required for<lb/> efficient HIV-1
			infection. To a large extent this has been facilitated by the<lb/> use of a potent and
			specific small molecule inhibitor of ATM, KU-55933,<lb/> and by using this inhibitor in
			studies that extend beyond vector-based<lb/> transduction experiments, to include HIV-1
			replication assays and clini-<lb/>cal HIV-1 isolate infections of primary
			cells.<lb/></p>

		<p>The current confusion surrounding the role of DNA damage response<lb/> proteins in
			retroviral infections <ref type="biblio">7,8,23,24</ref> probably stems from the use
			of<lb/> non-specific inhibitors such as caffeine and wortmannin, the ectopic<lb/>
			overexpression of dominant-negative proteins, over-interpretation of<lb/> short
			interfering RNA (siRNA) knockdown data and the use of very dif-<lb/>ferent vector-based
			transduction assay conditions. For example, caffeine<lb/> has been implicated in
			inhibiting ATM, ATR <ref type="biblio">25</ref> , DNA-PK <ref type="biblio">26</ref> ,
			CHK2 <ref type="biblio">(ref.<lb/> 27)</ref>, PI(3)K <ref type="biblio">28</ref> and
			mTOR <ref type="biblio">29</ref> , meaning that it cannot be used to conclusively<lb/>
			differentiate between these kinase activities in vivo <ref type="biblio">26</ref> .
			Wortmannin is not<lb/> much more specific, with inhibitory activity primarily against
			PI(3)K but<lb/> with the compound also effectively inhibiting DNA-PK, ATM and ATR<lb/>
			at micromolar concentrations <ref type="biblio">29</ref> . In published siRNA knockdown
			experi-<lb/>ments, non-specific control siRNA molecules alone have been shown<lb/> to
			activate retroviral transduction readouts <ref type="biblio">23</ref> , whereas
			dominant-nega-<lb/>tive kinase dead ATR constructs <ref type="biblio">30</ref> — used in
			an attempt to differentiate<lb/> between ATM and ATR contributions to retroviral
			transduction <ref type="biblio">7</ref> — have<lb/> previously been demonstrated to also
			have inhibitory effects on ionizing-<lb/>radiation-induced, ATM-driven phosphorylation
			events <ref type="biblio">31</ref> . With such a<lb/> combination of different methods
			used, it is perhaps not surprising that<lb/> conflicting interpretations have
			arisen.<lb/></p>

		<p>The earliest published study of the involvement of ATM in retro-<lb/>viral transduction
			came to the conclusion that this kinase only had<lb/> a minor role in retroviral
			infections <ref type="biblio">7</ref> . In contrast, our findings dem-<lb/>onstrate that
			ATM has a significant role in HIV-1 replication and the<lb/> data obtained with the ATM
			kinase inhibitor KU-55933 provide proof<lb/> of concept that this approach may have
			potential as an anti-retroviral<lb/> therapy. Possible reasons for these different
			conclusions may hinge<lb/> on Skalka and colleagues&apos; use of wortmannin along with
			their use of<lb/> different control cell lines. What is less easy to explain is why
			their<lb/> study reported no difference in transduction efficiencies between the<lb/>
			ATM-deficient cell line AT22IJE-T and the matched ATM-reconsti-<lb/>tuted AT22IJE-T
			cells, whereas in this study a clear difference was<lb/> observed <ref type="figure"
				>(Fig. 1b)</ref>. One possible explanation may come from the fact<lb/> that ATM
			expression can be lost from the complimented cells and it<lb/> is important to check the
			phenotype of the reconstituted cells. In the<lb/> experiments presented in this study,
			both lines were assayed to deter-<lb/>mine their response to ionizing-radiation-induced
			DNA damage in<lb/> parallel with the HIV-1 transduction experiments, and the results
			for<lb/> both the radiation survival assays (see <ref type="figure">Supplementary
				Information,<lb/></ref></p>

		<figure type="table">Table 1 KU-55933 inhibits drug-resistant HIV-1 as well as wild
			type<lb/> Drug IC 50 (µM)<lb/> HIV RT mutant<lb/> KU-55933<lb/> Abacavir<lb/>
			Lamivudine<lb/> WT<lb/> 3.8<lb/> 2.9<lb/> 0.6<lb/> L74V/M184V<lb/> 5.2<lb/> 20<lb/>
			&gt;100<lb/> K65R/M184V<lb/> 4.9<lb/> 11.3<lb/> &gt;100<lb/> K65R/L74V/M184V<lb/>
			2.8<lb/> 11.0<lb/> &gt;100<lb/> Values given represent data from two independent
			experiments.<lb/></figure>

		<p><ref type="figure">Fig. S2</ref>) and the HIV-1 vector transduction assays <ref
				type="figure">(Fig. 1b)</ref> are<lb/> consistent. In addition, although the
			ATM-reconstituted AT22IJE-T<lb/> line is more resistant to ionizing radiation and more
			capable of sup-<lb/>porting HIV-1 vector transduction than the ATM-deficient
			AT22IJE-<lb/>T cells, they are not fully complemented for the AT defect <ref
				type="biblio">31</ref> .<lb/></p>

		<p>Our analyses of DNA damage response markers following HIV-1 infec-<lb/>tion also provide
			a number of important insights into the involvement of<lb/> ATM in this process. First,
			in response to retroviral infection, phosphor-<lb/>ylation of ATM Ser 1981, CHK2 Thr 68,
			p53 Ser 15 and NBS1 Ser 343 are<lb/> only observed in cells with functional ATM. Second,
			the results that we<lb/> obtained from infection with IN D64V mutant HIV-1 demonstrated
			reduced<lb/> levels of ATM substrate phosphorylation compared with infection with<lb/>
			integrase-proficient virus, suggesting that ATM signalling is in response<lb/> to the
			integrase-induced cleavage of host genomic DNA and not as a<lb/> consequence of the
			presence of double-stranded viral cDNA. Perhaps<lb/> the one exception to this is the
			autophosphorylation of ATM Ser 1981.<lb/> This phosphorylation event is a very sensitive
			marker of cellular stress<lb/> that is not absolutely dependent upon the generation of
			DNA DSBs and<lb/> can probably occur as a consequence of changes in chromatin structure
				<ref type="biblio">17</ref> .<lb/> The fact that our observations show HIV-1 IN D64V
			mutant virus failing<lb/> to generate an extended ATM signalling response (with CHK2 and
			p53<lb/> phosphorylation), suggests that either a critical threshold of DNA damage<lb/>
			has not been reached or that there are differences in composition between<lb/> the
			protein complexes (and therefore available substrates) associated with<lb/> linear viral
			cDNA ends and viral–host DNA integration intermediates.<lb/> Although not strictly a DNA
			double-strand break, the close proximity of<lb/> integrase-induced DNA breaks (see <ref
				type="figure">Supplementary Information, Fig. S3</ref>)<lb/> indicate that it may be
			recognized and dealt with as such. Notably, a recent<lb/> study by Skalka and colleagues
			described the phosphorylation and foci<lb/> formation of histone H2AX in response to
			retroviral infection <ref type="biblio">32</ref> . ATM is<lb/> thought to be the
			predominant kinase that phosphorylates H2AX Ser 139<lb/> in response to DNA DSBs and
			this process is thought to be ATR-inde-<lb/>pendent in non-replicating cells <ref
				type="biblio">33</ref> . These data are therefore consistent with<lb/> ATM
			activation at a time after infection during which we also observe<lb/> other
			ATM-dependent phosphorylation events <ref type="figure">(Fig. 2)</ref>.<lb/></p>

		<p>Activation of an ATM signalling cascade after integrase-induced<lb/> DNA damage is also
			consistent with a number of additional observa-<lb/>tions in this study, which provide
			an insight into the mechanism of<lb/> action of KU-55933 as an anti-retroviral agent.
			The infection with an<lb/> integrase-proficient HIV-1 vector induces cell death in
			ATM-deficient<lb/> cells <ref type="figure">(Fig. 3)</ref>, suggesting that ATM provides
			the host cell with protec-<lb/>tion against HIV-1 integrase-induced DNA damage.
			Consequently, in<lb/> the absence of functional ATM the host cell is more likely to die,
			thus<lb/> preventing the establishment of an infection. KU-55933 reduces the<lb/>
			efficiency of HIV-1 transduction down to a level that is seen for AT<lb/> cells, and
			ablates ATM-signalling as indicated by the lack of p53 Ser 15<lb/> phosphorylation <ref
				type="figure">(Fig. 4)</ref>, indicating that the effect of the compound in<lb/>
			these experiments is through the inhibition of ATM. KU-55933 does<lb/> not affect HIV-1
			reverse transcriptase activity, viral nuclear import or<lb/> the HIV-1 integrase
			protein&apos;s ability to catalyse the joining of viral cDNA<lb/> with the host genomic
			DNA <ref type="figure">(Fig. 5)</ref>. Also consistent with the inhibition<lb/> of
			post-integration repair by KU-55933, is the loss of integrated viral<lb/> DNA from the
			cell population after 48 h, consistent with the induction<lb/> of host cell death.
			Further evidence that the compound is not inhibiting<lb/> a retroviral target such as
			reverse transcriptase or protease activity in<lb/> replication assays, is the lack of
			any antagonism in drug combination<lb/> studies <ref type="figure">(Fig. 6)</ref> and
			the ability of KU-55933 to work just as effectively<lb/> against drug-resistant HIV-1
				<ref type="table">(Table 1)</ref>.<lb/></p>

		<p>The data presented therefore suggest that small molecule inhibitors of<lb/> ATM have the
			potential to suppress HIV-1 infections. Although other DNA<lb/> damage response proteins
			have been implicated in modulating retroviral<lb/> infections <ref type="biblio"
				>3–8,34–38</ref> many of these may not be suitable targets for anti-retroviral<lb/>
			therapy. For example, ATR seems to be essential for normal cell function <ref
				type="biblio">39,40</ref><lb/> and it is likely that if specific ATR kinase
			inhibitors could be identified,<lb/> they would result in cellular lethality. Other
			targets may be implicated from<lb/> genetic data but their enzymatic activities may not
			be involved. For example,<lb/> both DNA-PK and PARP-1 have been shown to positively
			influence retrovi-<lb/>ral transduction <ref type="biblio">3–6,34,35</ref> , but our
			potent and specific small molecule inhibitors<lb/> of DNA-PK <ref type="biblio">29</ref>
			and PARP <ref type="biblio">41</ref> have little or no effect on retroviral
			transduction<lb/> (data not shown). Ultimately, whichever cellular target is being
			considered<lb/> for anti-HIV-1 therapy, there will be associated challenges. Although
			not<lb/> essential for cell survival (AT patients generally live for 20–30 years),
			the<lb/> benefit-to-risk equation of targeting ATM for HIV-1 infections will have
			to<lb/> be carefully investigated. In spite of the challenges, perhaps intelligent
			and<lb/> intermittent use of inhibitors of cellular targets such as ATM will provide
			a<lb/> much needed additional weapon in the fight against HIV-1 infections.<lb/></p>

		<figure>Untreated<lb/> 10 µM KU-55933<lb/> 1 µM zidovudine<lb/> a<lb/> b<lb/> c<lb/>
			Untreated<lb/> 10 µM KU-55933<lb/> 1 µM zidovudine<lb/> Untreated<lb/> 10 µM
			KU-55933<lb/> 1 µM zidovudine<lb/> Untreated<lb/> 10 µM KU-55933<lb/> 1 µM
			zidovudine<lb/> Hours post-infection<lb/> Hours post-infection<lb/> Hours
			post-infection<lb/> Late-RT DNA<lb/> 2-LTR DNA<lb/> Integrated DNA<lb/> 0<lb/> 0 6 12 18
			24 30 36 42 48 54<lb/> 0.5<lb/> 1.0<lb/> 1.5<lb/> 2.0<lb/> 2.5<lb/> 3.0<lb/> Hours
			post-infection<lb/> 0 6 12 18 24 30 36 42 48 54<lb/> Hours post-infection<lb/> 0 6 12 18
			24 30 36 42 48 54<lb/> Hours post-infection<lb/> 0<lb/> 0.2<lb/> 0.4<lb/> 0.6<lb/>
			0.8<lb/> 1.0<lb/> 1.2<lb/> 0<lb/> 0.5<lb/> 1.0<lb/> 1.5<lb/> 2.0<lb/> 2.5<lb/> 0 3 6 9
			24 48<lb/> 0 3 6 9 24 48<lb/> 0 3 6 9 24 48<lb/> Late-RT DNA (relative to GAPDH)<lb/>
			2-LTR DNA<lb/> (relative to GAPDH)<lb/> Integrated DNA<lb/> (relative to GAPDH)<lb/>
			Figure 5 Effect of KU-55933 on HIV-1 cDNA intermediate formation and<lb/> integration.
			(a–c) PCR of DNA extracted from HIV-1 IN + vector-transduced<lb/> 293 cells at
			increasing times after virus addition. The amounts of HIV-1<lb/> late-RT cDNA (a), 2-LTR
			cDNA (b) and integrated proviral DNA (c) were<lb/> estimated from cells that were
			transduced in the absence (untreated; blue<lb/> squares) or presence of 10 µM KU-55933
			(pink crosses) or 1 µM zidovudine<lb/> (orange triangles). Left panels show gel images
			of PCRs and right panels<lb/> show the quantified, normalized data. All PCR
			quantification results are<lb/> expressed as a normalized ratio of HIV-1 DNA: GAPDH
			control DNA. Gel<lb/> images of GAPDH control PCRs are not shown.<lb/></figure>

		<head>METHODS<lb/></head>

		<p>Cell lines. The murine ES cell line J1 and its ATM −/− homozygous knockout<lb/>
			derivative <ref type="biblio">42</ref> were a kind gift from F. Alt and were grown in
			the absence of feeder<lb/> cells on gelatinized cell culture dishes in DMEM with 15% FBS
			and supplemented<lb/> with 500 units ml −1 LIF (Chemicon International, Temecula, CA).
			Human SV40-<lb/>transformed skin fibroblast cells 1BR.3.N, p53-positive U2OS cells and
			293 cells<lb/> were obtained from The European Collection of Cell Cultures (ECACC;
			Salisbury,<lb/> UK). SV40-transformed AT skin fibroblast cells AT5BIVA and
			EBV-transformed<lb/> GM14680 normal and GM01526 AT lymphoblastoid cells were all
			obtained from<lb/> Coriell Cell Repositories (Coriell Institute for Medical Research,
			NJ). ATM-com-<lb/>plimented AT22IJE-T pEBS7-YZ5 and vector-only control AT22IJE-T
			pEBS7<lb/> human fibroblast cells were a kind gift from Y. Shiloh and have been
			previously<lb/> described <ref type="biblio">13</ref> . Jurkat and C8166
			T-lymphoblastoid cells were grown as suspension<lb/> cultures in RPMI 1640 medium
			supplemented with 10% FBS.<lb/></p>

		<p>HIV-1 transduction assays. Single-round transduction assays using HIV-1 vectors<lb/> and
			based on colony formation and luciferase have been previously described <ref
				type="biblio">12</ref> .<lb/> Luciferase transduction assays done in the presence of
			drug were performed by<lb/> infecting cells with retroviral stocks for 1 h (multiplicity
			of infection (MOI) ~0.25)<lb/> before washing and replacing with fresh medium containing
			dilutions of KU-55933.<lb/> At least four replicate wells were performed for each drug
			dilution. Cells were incu-<lb/>bated for 72 h and then assayed for luciferase activity
			as previously described <ref type="biblio">12</ref> .<lb/></p>

		<p>HIV-1 replication p24 and XTT cytopathicity assays with recombinant HIV-<lb/>1 vectors.
			Recombinant HIV-1 RF and HIV-1 HXB2 wild-type viruses and deriva-<lb/>tives containing
			site-directed mutations in the reverse transcriptase gene were<lb/> obtained from the
			MRC AIDS Directed Reagent Programme (National Institute<lb/> Biological Standards and
			Control, Potters Bar, UK). C8166 T-cells were infected<lb/> with an HIV-1 strain for 1–2
			h before being washed and re-seeded into fresh<lb/> 96-well plates containing dilutions
			of KU-55933, abacavir or lamivudine. At<lb/> least three replicate wells were performed
			for each drug dilution. For HIV-1 p24<lb/> antigen assays, cells were incubated for 7
			days before cell-free supernatants were<lb/> collected and virus quantified by HIV-1 p24
			antigen enzyme-linked immuno-<lb/>sorbent assay (ELISA; Abbott Laboratories, Abbot Park,
			IL). The concentration<lb/> of drug that gave a 50% reduction in HIV-1 p24 when compared
			with untreated<lb/> control cells (IC 50 ) was calculated from sigmoidal plots of the
			amount of HIV-1<lb/> p24 versus drug concentration. For HIV-1 XTT (a tetrazolium salt)
			cytopathic-<lb/>ity assays, cells were incubated for 5–6 days and viable cells remaining
			were<lb/> quantified using the XTT assay <ref type="biblio">43</ref> . In these
			cytopathicity assays, the number of<lb/> viable cells remaining is inversely
			proportional to the level of HIV-1 infection.<lb/> The concentration of drug that gave a
			50% inhibition of HIV-1 cytopathicity<lb/> relative to the untreated control cells (IC
			50 ) was calculated from sigmoidal plots<lb/> of cell viability versus drug
			concentration. For KU-55933 combination stud-<lb/>ies, C8166 cells were infected with
			wild-type HIV- 1 HXB2 for 1–2 h before being<lb/> washed and seeded into 96-well plates
			containing increasing concentrations of<lb/> KU-55933. Different concentrations of
			lamivudine were then added and cells<lb/> were incubated for 7 days. IC 50 values for
			inhibitors alone or in combination<lb/> were determined using XTT cytopathicity assays
			and plotted against each other<lb/> to give isobolograms <ref type="biblio">22</ref> . A
			curve that essentially follows the line between the IC 50<lb/> values of the two drugs
			would indicate additivity, whereas a convex curve would<lb/> indicate drug antagonism,
			and a concave curve drug synergy.<lb/></p>

		<p>HIV-1 replication assays with HIV-1 clinical isolates. Replication assays<lb/> using
			HIV-1 clinical isolates were performed on 3-day IL-2-stimulated<lb/> human PBMCs
			isolated from buffy coat residues (Lymphoprep; Axis-Shield,<lb/> Oslo, Norway) of blood
			from healthy volunteers (National Blood service,<lb/> UK). PBMCs were infected for 1–2 h
			with either wild-type clinical isolates of<lb/> HIV-1 CC15 , HIV-1 CC1 and HIV-1 105A ,
			lamivudine-resistant HIV-1 105F , or zido-<lb/>vudine-resistant HIV-1 CC18 clinical
			HIV-1 virus isolates <ref type="biblio">44 , before washing and<lb/> re-seeding into
				fresh 48-well plates that contained dilutions of KU-55933. At<lb/> le</ref>ast three
			replicate wells were performed for each drug dilution. Cells were<lb/> incubated for 7
			days and the amount of HIV-1 virus in cell-free supernatants<lb/> was quantified by
			HIV-1 p24 antigen ELISA as previously described. PBMCs<lb/> and clinical HIV-1 strains
			were isolated and phenotyped by MRC-Technology<lb/> (London, UK) with all of the
			relevant ethical approvals.<lb/></p>

		<p>Drug cytotoxicity assays. For all cell-based drug inhibition studies, cytotoxicity<lb/>
			assays were performed in parallel to monitor effects of drug alone on cell growth.<lb/>
			Cytotoxicity/cell viability assays were performed using at least three replicate
			wells<lb/> by either XTT <ref type="biblio">43</ref> or MTS (CellTiter96-aqueous
			solution; Promega, Madison, WI)<lb/> viability assays. The concentration of drug that
			gave 50% reduction in cell viability<lb/> when compared with untreated control cells (CC
			50 ) was calculated from sigmoidal<lb/> plots of cell viability versus drug
			concentration.<lb/></p>

		<p>Cytopathicity assays. Mouse ES cells were plated on gelatinized 96-well plates<lb/> and
			either mock infected (cell culture supernatants without virus) or infected with<lb/>
			HIV-1 IN + , HIV-1 IN D64V or heat-inactivated HIV-1 IN + luciferase virus vector<lb/>
			stocks at an MOI~10. Heat-inactivated HIV-1 IN + stocks were incubated at 60 o C<lb/>
			for 10 min. At increasing times after virus addition, cells were collected and
			stained<lb/> with annexin-V–EGFP (Clontech, Palo Alto, CA) according to the
			manufacturers&apos;<lb/> instructions. The percentage of annexin-V-positive cells was
			determined by flow<lb/> cytometry analysis using BD FACScalibur and CellQuest
			software.<lb/></p>

		<p>HIV-1 induced ATM activation. U2OS, GM14680 or GM01526 AT cells<lb/> (1.5 × 10 6 in each
			case) were transduced with HIV-1 IN + or IN D64V mutant HIV-<lb/>1 luciferase virus
			stocks (MOI~5) for 1 h before changing to fresh medium. At<lb/> increasing times after
			virus exposure, cells were washed twice in PBS, collected,<lb/> and split at a ratio of
			2:3 for preparation of protein extracts and 1:3 for DNA<lb/> extraction. Whole-cell
			protein lysates were prepared by resuspending cells in<lb/> either SDS-loading buffer
			(50 mM Tris pH 6.8, 2% SDS, 10% glycerol, 0.5 mM<lb/></p>

		<figure>a<lb/> IC 50<lb/> CC 50<lb/> 2.3 µM<lb/> 0.9 µM<lb/> 0.2 µM<lb/> KU-55933<lb/>
			Abacavir<lb/> Lamivudine<lb/> 0<lb/> 20<lb/> 40<lb/> 60<lb/> 80<lb/> 100<lb/> 120<lb/>
			Log drug concentration [µM]<lb/> 1<lb/> 10<lb/> 10 −1<lb/> 10 −2<lb/> 10 −3<lb/> 10
			−4<lb/> 100<lb/> 0<lb/> 0.4<lb/> 0.8<lb/> 1.2<lb/> 1.6<lb/> 2.0<lb/> KU-55933 ( µM)<lb/>
			Fixed [KU-55933]<lb/> Fixed [lamivudine]<lb/> IC 50<lb/> b<lb/> HIV-1 replication<lb/>
			(level of p24 as percentage control)<lb/> &gt; 30 µM<lb/> &gt; 30 µM<lb/> &gt; 30
			µM<lb/> Lamivudine (µM)<lb/> 0<lb/> 2<lb/> 4<lb/> 6<lb/> 8<lb/> 1 0<lb/> 1 2<lb/> Figure
			6 Inhibition of HIV-1 replication by KU-55933. (a) HIV-1 replication<lb/> assays for
			C8166 cells infected with HIV-1 HXB2 wild-type virus in the presence<lb/> of KU-55933
			(blue diamonds) or the reverse transcriptase inhibitors abacavir<lb/> (pink crosses) and
			lamivudine (orange circles). The mean percentage of HIV-1<lb/> p24 antigen compared with
			untreated control cells from at least two replicate<lb/> infections along with compound
			IC 50 and CC 50 values are shown. (b) Inhibition<lb/> of HIV-1 replication by KU-55933
			used in combination with lamivudine.<lb/> C8166 cells were infected with wild-type HIV-1
			HXB2 in the presence of<lb/> increasing concentrations of inhibitor alone and in
			combination with the other<lb/> inhibitor. IC 50 values were determined for each
			inhibitor combination by XTT<lb/> cytopathicity assays from at least two replicate
			infections. The data are shown<lb/> as IC 50 isobologram plots, and inhibitor
			combinations that fall below the<lb/> dotted line drawn between the IC 50 values for the
			individual inhibitors (pink<lb/> triangles) indicate synergistic effects.<lb/></figure>

		<p>MgOAc and 1 mM DTT) or high-salt buffer (20 mM HEPES pH 7.2, 450 mM<lb/> NaCl, 10%
			glycerol, 0.2 mM EDTA, 1 mM DTT, 5 mM NaF, 1 mM Na 3 VO 4 and<lb/> protease inhibitor
			cocktail) followed by freeze thaw extraction. Proteins were<lb/> separated on 6% or 10%
			SDS–PAGE gels, blotted on PVDF membranes and<lb/> probed with phosphospecific rabbit
			polyclonal antibodies against ATM Ser 1981<lb/> (Rockland, Gilbertsville, PA), CHK2 Thr
			68, p53 Ser 15 or NBS1 Ser 343 (Cell<lb/> Signaling Technology, Beverly, MA).
			Immunoblots were also probed with<lb/> antibodies against ATM (5C2; Abcam, Cambridge,
			UK), CHK2 (NT2; Prosci,<lb/> Anaheim, CA), p53 (DO-1; Santa Cruz Biotech, Santa Cruz,
			CA), NBS1 (NB100-<lb/>143; Abcam), ATR (N-19; Santa Cruz Biotech), DNA-PK (Ab-2;
			Labvision,<lb/> Fremont, CA) or β-actin antibody (Sigma, St Louis, MO). As a control for
			viral<lb/> entry and reverse transcription of IN D64V mutant virus vectors, PCRs for
			the<lb/> presence of HIV-1 vector DNA were performed on total DNA that had been<lb/>
			extracted from each sample. The amount of HIV-1 viral cDNA (late-RT and<lb/> 2-LTR) in
			IN + and IN D64V mutant transduced cells was found to be similar <ref type="biblio"
				>12</ref> (see<lb/> also <ref type="figure">Supplementary Information, Fig.
			S4</ref>). For radiation-treated samples, sub-<lb/>confluent cells were exposed to
			either 2 Gy ionizing radiation using a Faxitron<lb/> Corporation (Wheeling, IL) X-ray
			cabinet, or 20 J m −2 ultraviolet radiation using<lb/> a Stratagene (La Jolla, CA)
			Stratalinker.<lb/></p>

		<p>PCR analysis of HIV-1 DNA. PCR analysis of HIV-1 late-RT, 2-LTR and
			glyc-<lb/>eraldehyde-3-phosphate dehydrogenase (GAPDH) DNA from HIV-1
			vec-<lb/>tor-transduced 293 cells has been described previously <ref type="biblio"
				>12</ref> . Integrated proviral<lb/> HIV-1 DNA was quantified using a nested HIV-1
			LTR-Alu PCR. First-round<lb/> LTR-Alu PCRs were performed using 50 ng genomic DNA with
			HIV-1 LTR<lb/> primer MH535 and Alu primer SB704 <ref type="biblio">(ref. 45)</ref>. We
			then amplified and quanti-<lb/>fied 1:1,000 of first-round PCR product using HIV-1 LTR
			nested primer pairs<lb/> LTR6 and LTR9 <ref type="biblio">(ref. 46)</ref>. At this
			dilution, unintegrated HIV-1 LTR DNA car-<lb/>ryover was not detectable in PCR control
			reactions or in HIV-1 IN D64V infected<lb/> cells (see <ref type="figure">Supplementary
				Information, Fig. S4</ref>), ensuring that all detectable LTR<lb/> amplification
			corresponds to integrated proviral DNA only. All PCRs were<lb/> limited in cycle number
			to ensure linearity of amplification.<lb/></p>

	</text>
</tei>