Reactivation of HIV-1 from Latency by an Ingenol Derivative from Euphorbia Kansui

Pengfei Wang, Panpan Lu, Xiying Qu, Yinzhong Shen, Hanxian Zeng, Xiaoli Zhu, Yuqi Zhu, Xian Li, Hao Wu, Jianqing Xu, Hongzhou Lu, Zhongjun Ma, Huanzhang Zhu, Pengfei Wang, Panpan Lu, Xiying Qu, Yinzhong Shen, Hanxian Zeng, Xiaoli Zhu, Yuqi Zhu, Xian Li, Hao Wu, Jianqing Xu, Hongzhou Lu, Zhongjun Ma, Huanzhang Zhu

Abstract

Cells harboring latent HIV-1 pose a major obstacle to eradication of the virus. The 'shock and kill' strategy has been broadly explored to purge the latent reservoir; however, none of the current latency-reversing agents (LRAs) can safely and effectively activate the latent virus in patients. In this study, we report an ingenol derivative called EK-16A, isolated from the traditional Chinese medicinal herb Euphorbia kansui, which displays great potential in reactivating latent HIV-1. A comparison of the doses used to measure the potency indicated EK-16A to be 200-fold more potent than prostratin in reactivating HIV-1 from latently infected cell lines. EK-16A also outperformed prostratin in ex vivo studies on cells from HIV-1-infected individuals, while maintaining minimal cytotoxicity effects on cell viability and T cell activation. Furthermore, EK-16A exhibited synergy with other LRAs in reactivating latent HIV-1. Mechanistic studies indicated EK-16A to be a PKCγ activator, which promoted both HIV-1 transcription initiation by NF-κB and elongation by P-TEFb signal pathways. Further investigations aimed to add this compound to the therapeutic arsenal for HIV-1 eradication are in the pipeline.

Conflict of interest statement

The authors declare that they have no competing interests.

Figures

Figure 1
Figure 1
Reactivation of latent HIV-1 by compounds extracted from Euphorbia kansui. (a) C11 cells were treated with various compounds extracted from Euphorbia kansui at three concentrations (0.1 μg/ml, 1 μg/ml and 10 μg/ml). At 48 h post-treatment, the percentage of GFP-positive cells was measured by flow cytometry, which represented the level of HIV-1 transcription. Data show the highest GFP induction of each compound at the three concentrations tested. (b) C11 cells were treated with the 8 selected compounds at the indicated concentrations for 48 h and the percentage of GFP-positive cells was measured. Data show the means ± standard deviations (s.d.) of three independent experiments.
Figure 2
Figure 2
Reactivation of latent HIV-1 in latently infected cells by EK-16A and prostratin. (a) The structure of EK-16A and prostratin. (b) C11 cells and (c) J-Lat 10.6 cells were treated with EK-16A or prostratin at the indicated concentrations. At 48 h post-treatment, the percentage of GFP-positive cells was measured by flow cytometry and dose-dependent curves were shown. Data show the means ± s.d. of three independent experiments.
Figure 3
Figure 3
Effects of EK-16A on cell viability. C11 cells (a) and J-Lat 10.6 cells (b) were treated with EK-16A or prostratin at the indicated concentration for 72 h and then cell viability was measured by CCK-8 kit (Dojindo). The division of OD450 between treated and control groups indicate the percentage of cell viability. Data show the means ± s.d. in three independent experiments.
Figure 4
Figure 4
EK-16A synergizes with other LRAs in reactivating latent HIV-1. J-Lat 10.6 (a) and J-Lat 6.3 cells (b) were treated with EK-16A at the indicated concentrations alone or in combination with prostratin (0.5 μM), 5-Aza (1 μM), JQ1 (0.5 μM), I-Bet151 (0.5 μM), romidepsin (5 nM), or vorinostat (0.5 μM) for 24 h and and the percentage of GFP-positive cells was measured. (c) The Bliss independence model was utilized for calculation of synergy for LRA combinations (See materials and methods). Dotted horizontal line signifies pure additive effect (Δfaxy = 0). Synergy is defined as Δfaxy > 0 while Δfaxy < 0 indicates antagonism. Data show the means ± s.d. of three independent experiments.
Figure 5
Figure 5
EK-16A reactivates HIV-1 from infected resting CD4+ T cells ex vivo. Resting CD4 T cells isolated from five cART-suppressed HIV-1-infected patients were treated with either EK-16A (0.05 μM) or prostratin (1 μM) for 18 h. Intracellular HIV-1 mRNA levels were detected by qRT-PCR and presented as fold induction relative to mock-treated control. **p < 0.01, two-tailed unpaired Student t test, n = 5.
Figure 6
Figure 6
The effects of EK-16A on T cell activation. (a) The effect of EK-16A and prostratin on the expression of CD25 and CD69. Human CD4+ T cells were treated with either EK-16A (0.05 μM) or prostratin (1 μM) for 24 h and the expression of CD25 and CD69 was detected by flow cytometry using antibodies against CD25 and CD69. The percentage of CD69-positive cells (b) and CD25-positive cells (c) was calculated. Data show the means ± s.d. of three independent experiments. (d) Human CD4+ T cells were treated with EK-16A (0.05 μM) or prostratin (1 μM) for 24 h and cell culture supernatant was assayed for cytokine concentrations (pg/ml) by ELISA. Data are representative of three independent experiments.
Figure 7
Figure 7
EK-16A activates HIV-1 mainly through PKCγ. (a) C11 cells were treated with 0.05 μM EK-16A alone (−), or pretreated with Gö 6983 (1 μM), Gö 6976 (1 μM), GF 109203X (1 μM), PKCθ/δ inhibitor (10 μM) or Rottlerin (1 μM) for 3 h, then treated with EK-16A for 48 h before percentage of GFP-positive cells was assessed by flow cytometry. (b) TZM-bl cells were untransfected (−) or transfected with negative control shRNA plasmid (sh-NC) or shRNA plasmids targeting different sequences of PKCγ for 24 h, then treated with 0.05 μM EK-16A for 48 h before firefly luciferase activity was measured. (c) C11 cells were nucleofected with indicated shRNA plasmids for 24 h, then treated with 0.05 μM EK-16A for 72 h before percentage of GFP-positive cells in the shRNA-expressing cells (RFP-positive) was measured by flow cytometry; (d) shows the corresponding data at the different indicated time points. All data represent the mean ± s.d. of three independent experiments.
Figure 8
Figure 8
EK-16A activates HIV-1 by up-regulation of NF-κB and P-TEFb. C11 cells were stimulated with prostratin (1 μM) or ΕΚ-16Α (0.05 μM) for 5, 10 or 30 min, and (a) total cell lysates were probed for the expression of total IκBα or phosphorylated IκBα (p-IκBα) in immunoblots. α-Tubulin expression was used as a control for protein loading. (b) Nuclear extracts were analyzed using antibodies against p65. TATA-binding protein (TBP) expression was used as a control for protein loading. (c) C11 cells were mock-treated or stimulated with prostratin (1 μM) or ΕΚ-16Α (0.05 μM) for 30 min. ChIP assays were performed using anti-p65 antibody and PCR primers for the LTR promoter. The percentage of input for each immunoprecipitation was calculated and the fold occupancy of p65 relative to mock-treatment is shown. Each ChIP experiment was repeated three times to confirm reproducibility of results. (d) C11 cells were stimulated with 0.05 μM EK-16A for 0.5, 1 or 2 h, and total cell lysates were probed for the expression of total CDK9, phosphorylated CDK9 (p-CDK9) or Cyclin T1 in immunoblots. α-Tubulin expression was used as a control for protein loading.
Figure 9
Figure 9
Diagram depicting EK-16A-mediated HIV-1 reactivation by both NF-κB and P-TEFb signaling pathways.

References

    1. Hammer SM, et al. A controlled trial of two nucleoside analogues plus indinavir in persons with human immunodeficiency virus infection and CD4 cell counts of 200 per cubic millimeter or less. AIDS Clinical Trials Group 320 Study Team. N. Engl. J. Med. 1997;337:725–733. doi: 10.1056/NEJM199709113371101.
    1. Gulick RM, et al. Treatment with indinavir, zidovudine, and lamivudine in adults with human immunodeficiency virus infection and prior antiretroviral therapy. N. Engl. J. Med. 1997;337:734–739. doi: 10.1056/NEJM199709113371102.
    1. Perelson AS, et al. Decay characteristics of HIV-1-infected compartments during combination therapy. Nature. 1997;387:188–191. doi: 10.1038/387188a0.
    1. Davey RT, Jr., et al. HIV-1 and T cell dynamics after interruption of highly active antiretroviral therapy (HAART) in patients with a history of sustained viral suppression. Proc. Natl. Acad. Sci. USA. 1999;96:15109–15114. doi: 10.1073/pnas.96.26.15109.
    1. Torres RA, Lewis W. Aging and HIV/AIDS: pathogenetic role of therapeutic side effects. Lab. Investig. 2014;94:120–128. doi: 10.1038/labinvest.2013.142.
    1. Hearps AC, Martin GE, Rajasuriar R, Crowe SM. Inflammatory co-morbidities in HIV+ individuals: learning lessons from healthy ageing. Curr. HIV/AIDS Rep. 2014;11:20–34. doi: 10.1007/s11904-013-0190-8.
    1. Samaras K. Prevalence and pathogenesis of diabetes mellitus in HIV-1 infection treated with combined antiretroviral therapy. J. Acquir. Immune Defic. Syndr. 2009;50:499–505. doi: 10.1097/QAI.0b013e31819c291b.
    1. Peterlin BM, Trono D. Hide, shield and strike back: how HIV-infected cells avoid immune eradication. Nat. Rev. Immunol. 2003;3:97–107. doi: 10.1038/nri998.
    1. Archin NM, et al. Eradicating HIV-1 infection: seeking to clear a persistent pathogen. Nat. Rev. Microbiol. 2014;12:750–764. doi: 10.1038/nrmicro3352.
    1. Battistini A, Sgarbanti M. HIV-1 latency: an update of molecular mechanisms and therapeutic strategies. Viruses. 2014;6:1715–1758. doi: 10.3390/v6041715.
    1. Ruelas DS, Greene WC. An integrated overview of HIV-1 latency. Cell. 2013;155:519–529. doi: 10.1016/j.cell.2013.09.044.
    1. Delagreverie HM, et al. Ongoing Clinical Trials of Human Immunodeficiency Virus Latency-Reversing and Immunomodulatory Agents. Open Forum Infect Dis. 2016;3:ofw189. doi: 10.1093/ofid/ofw189.
    1. Dahabieh MS, Battivelli E, Verdin E. Understanding HIV latency: the road to an HIV cure. Annu. Rev. Med. 2015;66:407–421. doi: 10.1146/annurev-med-092112-152941.
    1. Datta PK, et al. HIV-1 Latency and Eradication: Past, Present and Future. Curr. HIV Res. 2016;14:431–441. doi: 10.2174/1570162X14666160324125536.
    1. Shang HT, et al. Progress and challenges in the use of latent HIV-1 reactivating agents. Acta Pharmacol. Sin. 2015;36:908–916. doi: 10.1038/aps.2015.22.
    1. Margolis, D. M., Garcia, J. V., Hazuda, D. J. & Haynes, B. F. Latency reversal and viral clearance to cure HIV-1. Science353, aaf6517 (2016).
    1. Kulkosky J, et al. Intensification and stimulation therapy for human immunodeficiency virus type 1 reservoirs in infected persons receiving virally suppressive highly active antiretroviral therapy. J. Infect. Dis. 2002;186:1403–1411. doi: 10.1086/344357.
    1. Prins JM, et al. Immuno-activation with anti-CD3 and recombinant human IL-2 in HIV-1-infected patients on potent antiretroviral therapy. AIDS. 1999;13:2405–2410. doi: 10.1097/00002030-199912030-00012.
    1. Manson McManamy ME, Hakre S, Verdin EM, Margolis DM. Therapy for latent HIV-1 infection: the role of histone deacetylase inhibitors. Antivir. Chem. Chemother. 2014;23:145–149. doi: 10.3851/IMP2551.
    1. Lehrman G, et al. Depletion of latent HIV-1 infection in vivo: a proof-of-concept study. Lancet. 2005;366:549–555. doi: 10.1016/S0140-6736(05)67098-5.
    1. Siliciano JD, et al. Stability of the latent reservoir for HIV-1 in patients receiving valproic acid. J. Infect. Dis. 2007;195:833–836. doi: 10.1086/511823.
    1. Archin NM, et al. Valproic acid without intensified antiviral therapy has limited impact on persistent HIV infection of resting CD4+ T cells. AIDS. 2008;22:1131–1135. doi: 10.1097/QAD.0b013e3282fd6df4.
    1. Sagot-Lerolle N, et al. Prolonged valproic acid treatment does not reduce the size of latent HIV reservoir. AIDS. 2008;22:1125–1129. doi: 10.1097/QAD.0b013e3282fd6ddc.
    1. Archin NM, et al. Antiretroviral intensification and valproic acid lack sustained effect on residual HIV-1 viremia or resting CD4+ cell infection. PLoS One. 2010;5:e9390. doi: 10.1371/journal.pone.0009390.
    1. Routy JP, et al. Valproic acid in association with highly active antiretroviral therapy for reducing systemic HIV-1 reservoirs: results from a multicentre randomized clinical study. HIV med. 2012;13:291–296. doi: 10.1111/j.1468-1293.2011.00975.x.
    1. Archin NM, et al. HIV-1 expression within resting CD4+ T cells after multiple doses of vorinostat. J. Infect. Dis. 2014;210:728–735. doi: 10.1093/infdis/jiu155.
    1. Archin NM, et al. Administration of vorinostat disrupts HIV-1 latency in patients on antiretroviral therapy. Nature. 2012;487:482–485. doi: 10.1038/nature11286.
    1. Elliott JH, et al. Activation of HIV transcription with short-course vorinostat in HIV-infected patients on suppressive antiretroviral therapy. PLoS Pathog. 2014;10:e1004473. doi: 10.1371/journal.ppat.1004473.
    1. Rasmussen TA, et al. Panobinostat, a histone deacetylase inhibitor, for latent-virus reactivation in HIV-infected patients on suppressive antiretroviral therapy: a phase 1/2, single group, clinical trial. Lancet HIV. 2014;1:e13–21. doi: 10.1016/S2352-3018(14)70014-1.
    1. Wei DG, et al. Histone deacetylase inhibitor romidepsin induces HIV expression in CD4 T cells from patients on suppressive antiretroviral therapy at concentrations achieved by clinical dosing. PLoS Pathog. 2014;10:e1004071. doi: 10.1371/journal.ppat.1004071.
    1. Spivak AM, Planelles V. HIV-1 Eradication: Early Trials (and Tribulations) Trends Mol. Med. 2016;22:10–27. doi: 10.1016/j.molmed.2015.11.004.
    1. Darcis G, Van Driessche B, Van Lint C. Preclinical shock strategies to reactivate latent HIV-1: an update. Curr. Opin. HIV AIDS. 2016;11:388–393. doi: 10.1097/COH.0000000000000288.
    1. Williams SA, et al. Prostratin antagonizes HIV latency by activating NF-kappaB. J. Biol. Chem. 2004;279:42008–42017. doi: 10.1074/jbc.M402124200.
    1. Sung TL, Rice AP. Effects of prostratin on Cyclin T1/P-TEFb function and the gene expression profile in primary resting CD4+ T cells. Retrovirology. 2006;3:66. doi: 10.1186/1742-4690-3-66.
    1. Perez M, et al. Bryostatin-1 synergizes with histone deacetylase inhibitors to reactivate HIV-1 from latency. Curr. HIV Res. 2010;8:418–429. doi: 10.2174/157016210793499312.
    1. Diaz L, et al. Bryostatin activates HIV-1 latent expression in human astrocytes through a PKC and NF-kB-dependent mechanism. Sci. Rep. 2015;5:12442. doi: 10.1038/srep12442.
    1. Pandelo Jose D, et al. Reactivation of latent HIV-1 by new semi-synthetic ingenol esters. Virology. 2014;462–463:328–339. doi: 10.1016/j.virol.2014.05.033.
    1. Jiang G, et al. Reactivation of HIV latency by a newly modified Ingenol derivative via protein kinase Cdelta-NF-kappaB signaling. AIDS. 2014;28:1555–1566. doi: 10.1097/QAD.0000000000000289.
    1. Jiang G, et al. Synergistic Reactivation of Latent HIV Expression by Ingenol-3-Angelate, PEP005, Targeted NF-kB Signaling in Combination with JQ1 Induced p-TEFb Activation. PLoS Pathog. 2015;11:e1005066. doi: 10.1371/journal.ppat.1005066.
    1. Li Z, Guo J, Wu Y, Zhou Q. The BET bromodomain inhibitor JQ1 activates HIV latency through antagonizing Brd4 inhibition of Tat-transactivation. Nucleic Acids Res. 2013;41:277–287. doi: 10.1093/nar/gks976.
    1. Lu P, et al. The BET inhibitor OTX015 reactivates latent HIV-1 through P-TEFb. Sci. Rep. 2016;6:24100. doi: 10.1038/srep24100.
    1. Xing S, et al. Disulfiram reactivates latent HIV-1 in a Bcl-2-transduced primary CD4+ T cell model without inducing global T cell activation. J. Virol. 2011;85:6060–6064. doi: 10.1128/JVI.02033-10.
    1. Doyon G, Zerbato J, Mellors JW, Sluis-Cremer N. Disulfiram reactivates latent HIV-1 expression through depletion of the phosphatase and tensin homolog. AIDS. 2013;27:F7–F11. doi: 10.1097/QAD.0b013e3283570620.
    1. Li P, et al. Stimulating the RIG-I pathway to kill cells in the latent HIV reservoir following viral reactivation. Nat. Med. 2016;22:807–811. doi: 10.1038/nm.4124.
    1. Winckelmann AA, et al. Administration of a Toll-like receptor 9 agonist decreases the proviral reservoir in virologically suppressed HIV-infected patients. PLoS One. 2013;8:e62074. doi: 10.1371/journal.pone.0062074.
    1. Novis CL, et al. Reactivation of latent HIV-1 in central memory CD4(+) T cells through TLR-1/2 stimulation. Retrovirology. 2013;10:119. doi: 10.1186/1742-4690-10-119.
    1. Offersen R, et al. A Novel Toll-Like Receptor 9 Agonist, MGN1703, Enhances HIV-1 Transcription and NK Cell-Mediated Inhibition of HIV-1-Infected Autologous CD4+ T Cells. J. Virol. 2016;90:4441–4453. doi: 10.1128/JVI.00222-16.
    1. Bullen CK, et al. New ex vivo approaches distinguish effective and ineffective single agents for reversing HIV-1 latency in vivo. Nat. Med. 2014;20:425–429. doi: 10.1038/nm.3489.
    1. Kimata JT, Rice AP, Wang J. Challenges and strategies for the eradication of the HIV reservoir. Curr. Opin. Immunol. 2016;42:65–70. doi: 10.1016/j.coi.2016.05.015.
    1. Tu Y. The discovery of artemisinin (qinghaosu) and gifts from Chinese medicine. Nat. Med. 2011;17:1217–1220. doi: 10.1038/nm.2471.
    1. Hori T, et al. Procyanidin trimer C1 derived from Theobroma cacao reactivates latent human immunodeficiency virus type 1 provirus. Biochem. Biophys. Res. Commun. 2015;459:288–293. doi: 10.1016/j.bbrc.2015.02.102.
    1. Wang C, et al. A Natural Product from Polygonum cuspidatum Sieb. Et Zucc. Promotes Tat-Dependent HIV Latency Reversal through Triggering P-TEFb’s Release from 7SK snRNP. PLoS One. 2015;10:e0142739. doi: 10.1371/journal.pone.0142739.
    1. Yan X, et al. Processing of kansui roots stir-baked with vinegar reduces kansui-induced hepatocyte cytotoxicity by decreasing the contents of toxic terpenoids and regulating the cell apoptosis pathway. Molecules. 2014;19:7237–7254. doi: 10.3390/molecules19067237.
    1. Cary DC, Fujinaga K, Peterlin BM. Euphorbia Kansui Reactivates Latent HIV. PLoS One. 2016;11:e0168027. doi: 10.1371/journal.pone.0168027.
    1. Hou JJ, et al. A single, multi-faceted, enhanced strategy to quantify the chromatographically diverse constituents in the roots of Euphorbia kansui. J. Pharm. Biomed. Anal. 2014;88:321–330. doi: 10.1016/j.jpba.2013.08.049.
    1. Ding D, et al. Involvement of histone methyltransferase GLP in HIV-1 latency through catalysis of H3K9 dimethylation. Virology. 2013;440:182–189. doi: 10.1016/j.virol.2013.02.022.
    1. Qu X, et al. Zinc-finger-nucleases mediate specific and efficient excision of HIV-1 proviral DNA from infected and latently infected human T cells. Nucleic Acids Res. 2013;41:7771–7782. doi: 10.1093/nar/gkt571.
    1. Jiang G, Dandekar S. Targeting NF-kappaB signaling with protein kinase C agonists as an emerging strategy for combating HIV latency. AIDS Res. Hum. Retroviruses. 2015;31:4–12. doi: 10.1089/aid.2014.0199.
    1. Kulkosky J, et al. Prostratin: activation of latent HIV-1 expression suggests a potential inductive adjuvant therapy for HAART. Blood. 2001;98:3006–3015. doi: 10.1182/blood.V98.10.3006.
    1. Jordan A, Bisgrove D, Verdin E. HIV reproducibly establishes a latent infection after acute infection of T cells in vitro. EMBO J. 2003;22:1868–1877. doi: 10.1093/emboj/cdg188.
    1. Laird GM, et al. Ex vivo analysis identifies effective HIV-1 latency-reversing drug combinations. J. Clin. Investig. 2015;125:1901–1912. doi: 10.1172/JCI80142.
    1. Darcis G, et al. An In-Depth Comparison of Latency-Reversing Agent Combinations in Various In Vitro and Ex Vivo HIV-1 Latency Models Identified Bryostatin-1 + JQ1 and Ingenol-B + JQ1 to Potently Reactivate Viral Gene Expression. PLoS Pathog. 2015;11:e1005063. doi: 10.1371/journal.ppat.1005063.
    1. Fenaux P. Inhibitors of DNA methylation: beyond myelodysplastic syndromes. Nat. Clin. Pract. Oncol. 2005;2(Suppl 1):S36–44. doi: 10.1038/ncponc0351.
    1. Boehm D, et al. BET bromodomain-targeting compounds reactivate HIV from latency via a Tat-independent mechanism. Cell Cycle. 2013;12:452–462. doi: 10.4161/cc.23309.
    1. Blankson JN, Persaud D, Siliciano RF. The challenge of viral reservoirs in HIV-1 infection. Annu. Rev. Med. 2002;53:557–593. doi: 10.1146/annurev.med.53.082901.104024.
    1. Shan L, et al. A novel PCR assay for quantification of HIV-1 RNA. J. Virol. 2013;87:6521–6525. doi: 10.1128/JVI.00006-13.
    1. Gschwendt M, et al. Inhibition of protein kinase C mu by various inhibitors. Differentiation from protein kinase c isoenzymes. FEBS Lett. 1996;392:77–80. doi: 10.1016/0014-5793(96)00785-5.
    1. Martiny-Baron G, et al. Selective inhibition of protein kinase C isozymes by the indolocarbazole Go 6976. J. Biol. Chem. 1993;268:9194–9197.
    1. Cole DC, et al. Identification, characterization and initial hit-to-lead optimization of a series of 4-arylamino-3-pyridinecarbonitrile as protein kinase C theta (PKCtheta) inhibitors. J. Med. Chem. 2008;51:5958–5963. doi: 10.1021/jm800214a.
    1. Bedoya LM, et al. SJ23B, a jatrophane diterpene activates classical PKCs and displays strong activity against HIV in vitro. Biochem. Pharmacol. 2009;77:965–978. doi: 10.1016/j.bcp.2008.11.025.
    1. Oeckinghaus A, Ghosh S. The NF-kappaB family of transcription factors and its regulation. Cold Spring Harb. Perspect. Biol. 2009;1:a000034. doi: 10.1101/cshperspect.a000034.
    1. Mbonye UR, et al. Phosphorylation of CDK9 at Ser175 enhances HIV transcription and is a marker of activated P-TEFb in CD4(+) T lymphocytes. PLoS Pathog. 2013;9:e1003338. doi: 10.1371/journal.ppat.1003338.
    1. Rice AP. Cyclin-dependent kinases as therapeutic targets for HIV-1 infection. Expert Opin. Ther. Targets. 2016;20:1453–1461. doi: 10.1080/14728222.2016.1254619.
    1. Wang HY, et al. Bioactivity-guided isolation of antiproliferative diterpenoids from Euphorbia kansui. Phytother. Res. 2012;26:853–859. doi: 10.1002/ptr.3640.
    1. Zheng WF, Cui Z, Zhu Q. Cytotoxicity and antiviral activity of the compounds from Euphorbia kansui. Planta Med. 1998;64:754–756. doi: 10.1055/s-2006-957574.
    1. Zeng Y, et al. Screening of Epstein-Barr virus early antigen expression inducers from Chinese medicinal herbs and plants. Biomed. Environ. Sci. 1994;7:50–55.
    1. Nunomura S, Kitanaka S, Ra C. 3-O-(2,3-dimethylbutanoyl)-13-O-decanoylingenol from Euphorbia kansui suppresses IgE-mediated mast cell activation. Biol. Pharm. Bull. 2006;29:286–290. doi: 10.1248/bpb.29.286.
    1. Khiev P, et al. Ingenane-type diterpenes with a modulatory effect on IFN-gamma production from the roots of Euphorbia kansui. Arch. Pharm. Res. 2012;35:1553–1558. doi: 10.1007/s12272-012-0905-1.
    1. Wu TS, et al. Antitumor agents, 119. Kansuiphorins A and B, two novel antileukemic diterpene esters from Euphorbia kansui. J. Nat. Prod. 1991;54:823–829. doi: 10.1021/np50075a011.
    1. Wang LY, et al. Euphane and tirucallane triterpenes from the roots of Euphorbia kansui and their in vitro effects on the cell division of Xenopus. J. Nat. Prod. 2003;66:630–633. doi: 10.1021/np0205396.
    1. Doan HQ, Gulati N, Levis WR. Ingenol mebutate: potential for further development of cancer immunotherapy. J. Drugs Dermatol. 2012;11:1156–1157.
    1. Aditya S, Gupta S. Ingenol mebutate: A novel topical drug for actinic keratosis. Indian Dermatol. Online J. 2013;4:246–249.
    1. Fujiwara M, et al. Mechanism of selective inhibition of human immunodeficiency virus by ingenol triacetate. Antimicrob. Agents Chemother. 1996;40:271–273.
    1. Fujiwara M, et al. Upregulation of HIV-1 replication in chronically infected cells by ingenol derivatives. Arch. Virol. 1998;143:2003–2010. doi: 10.1007/s007050050436.
    1. Warrilow D, et al. HIV type 1 inhibition by protein kinase C modulatory compounds. AIDS Res. Hum. Retroviruses. 2006;22:854–864. doi: 10.1089/aid.2006.22.854.
    1. Abreu CM, et al. Dual role of novel ingenol derivatives from Euphorbia tirucalli in HIV replication: inhibition of de novo infection and activation of viral LTR. PLoS One. 2014;9:e97257. doi: 10.1371/journal.pone.0097257.
    1. Wender PA, et al. Analysis of the phorbol ester pharmacophore on protein kinase C as a guide to the rational design of new classes of analogs. Proc. Natl. Acad. Sci. USA. 1986;83:4214–4218. doi: 10.1073/pnas.83.12.4214.
    1. Newton AC. Protein kinase C: structure, function, and regulation. J. Biol. Chem. 1995;270:28495–28498. doi: 10.1074/jbc.270.48.28495.
    1. Kazi JU, Kabir NN, Ronnstrand L. Protein kinase C (PKC) as a drug target in chronic lymphocytic leukemia. Med. Oncol. 2013;30:757. doi: 10.1007/s12032-013-0757-7.
    1. Wu-Zhang AX, Newton AC. Protein kinase C pharmacology: refining the toolbox. Biochem. J. 2013;452:195–209. doi: 10.1042/BJ20130220.
    1. Soh JW, Weinstein IB. Roles of specific isoforms of protein kinase C in the transcriptional control of cyclin D1 and related genes. J. Biol. Chem. 2003;278:34709–34716. doi: 10.1074/jbc.M302016200.
    1. Zhong H, May MJ, Jimi E, Ghosh S. The phosphorylation status of nuclear NF-kappa B determines its association with CBP/p300 or HDAC-1. Mol. Cell. 2002;9:625–636. doi: 10.1016/S1097-2765(02)00477-X.
    1. Wei P, et al. A novel CDK9-associated C-type cyclin interacts directly with HIV-1 Tat and mediates its high-affinity, loop-specific binding to TAR RNA. Cell. 1998;92:451–462. doi: 10.1016/S0092-8674(00)80939-3.
    1. Kim YK, et al. Phosphorylation of the RNA polymerase II carboxyl-terminal domain by CDK9 is directly responsible for human immunodeficiency virus type 1 Tat-activated transcriptional elongation. Mol. Cell. Biol. 2002;22:4622–4637. doi: 10.1128/MCB.22.13.4622-4637.2002.
    1. Zhou Q, Li T, Price DH. RNA polymerase II elongation control. Annu. Rev. Biochem. 2012;81:119–143. doi: 10.1146/annurev-biochem-052610-095910.
    1. Sgarbanti M, Battistini A. Therapeutics for HIV-1 reactivation from latency. Curr. Opin. Virol. 2013;3:394–401. doi: 10.1016/j.coviro.2013.06.001.
    1. Spivak AM, et al. Janus kinase inhibition suppresses PKC-induced cytokine release without affecting HIV-1 latency reversal ex vivo. Retrovirology. 2016;13:88. doi: 10.1186/s12977-016-0319-0.
    1. Martin AR, et al. Rapamycin-mediated mTOR inhibition uncouples HIV-1 latency reversal from cytokine-associated toxicity. J. Clin. Invest. 2017;127:651–656. doi: 10.1172/JCI89552.
    1. Ying H, et al. Selective histonedeacetylase inhibitor M344 intervenes in HIV-1 latency through increasing histone acetylation and activation of NF-kappaB. PLoS One. 2012;7:e48832. doi: 10.1371/journal.pone.0048832.
    1. Wang P, et al. As2O3 synergistically reactivate latent HIV-1 by induction of NF-kappaB. Antiviral Res. 2013;100:688–697. doi: 10.1016/j.antiviral.2013.10.010.
    1. Wang P, et al. Specific reactivation of latent HIV-1 with designer zinc-finger transcription factors targeting the HIV-1 5′-LTR promoter. Gene Ther. 2014;21:490–495. doi: 10.1038/gt.2014.21.

Source: PubMed

3
Předplatit