AVR 3537

No. of Pages 16, Model 5G

11 November 2014 Antiviral Research xxx (2014) xxx–xxx 1

Contents lists available at ScienceDirect

Antiviral Research journal homepage: www.elsevier.com/locate/antiviral

2

Review

6 4 7

Nanotech-derived topical microbicides for HIV prevention: The road to clinical development

5 8

Q1

9 10 11 12 13 14 1 8 6 2 17 18 19 20 21 22 23 24 25 26 27

Javier Sánchez-Rodríguez a,b,1, Enrique Vacas-Córdoba a,b,1, Rafael Gómez c,b, F. Javier De La Mata c,b, Ma Ángeles Muñoz-Fernández a,b,⇑ a b

Q2

c

Laboratorio InmunoBiología Molecular, Hospital General Universitario Gregorio Marañón, Instituto de Investigación Sanitaria Gregorio Marañón, Madrid, Spain Networking Research Center on Bioengineering, Biomaterials and Nanomedicine (CIBER-BBN), Madrid, Spain Dendrimers for Biomedical Applications Group (BioInDen), University of Alcalá, Alcalá de Henares, Madrid, Spain

a r t i c l e

i n f o

Article history: Received 16 September 2014 Revised 20 October 2014 Accepted 29 October 2014 Available online xxxx Keywords: Microbicides HIV Nanotechnology Nanomedicine

a b s t r a c t More than three decades since its discovery, HIV infection remains one of the most aggressive epidemics worldwide, with more than 35 million people infected. In sub-Saharan Africa, heterosexual transmissions represent nearly 80% of new infections, with 50% of these occurring in women. In an effort to stop the dramatic spread of the HIV epidemic, new preventive treatments, such as microbicides, have been developed. Nanotechnology has revolutionized this field by designing and engineering novel highly effective nano-sized materials as microbicide candidates. This review illustrates the most recent advances in nanotech-derived HIV prevention strategies, as well as the main steps required to translate promising in vitro results into clinical trials. Ó 2014 Published by Elsevier B.V.

29 30 31 32 33 34 35 36 37 38 39 40

41 42 43 44 45 46 47 48 49 50 51

Contents 1. 2. 3. 4.

HIV prevention strategies: an urgent necessity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . History of the microbicide pipeline. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Nanotechnology: a new field in the design of microbicides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Moving towards pre-clinical development of nanotech-derived microbicides: the main stages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1. Identification of novel candidates: in vitro screening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1.1. Characteristics of ideal nanotech-derived microbicide candidates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1.2. Antiviral assays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1.3. Safety and biocompatibility of nanoparticles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1.4. HIV-transmission inhibition assays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

00 00 00 00 00 00 00 00 00

Abbreviations: AIDS, acquired immunodeficiency syndrome; ARV, antiretroviral; ASPIRE, a study to prevent infection with a ring for extended use; BLT, bone marrow liver thymic mice; CAP, cellulose acetate phthalate; CAPRISA, centre for the AIDS programme of research in South Africa; CBA, carbohydrate-binding agents; CCR5, C–C chemokine receptor type 5; CD34, cluster of differentiation 34 marker; CD4, cluster of differentiation 4 marker; CXCR4, C-X-C chemokine receptor type 4; DC-SIGN, dendritic cell-specific intercellular adhesion molecule-3-grabbing non-integrin; EC50, half maximal effective concentration; ELISA, enzyme-linked immunosorbent assay; HAART, highly active antiretroviral therapy; HIV, human immunodeficiency virus; HSV, herpes simplex virus; IL, interleukin; IPM, international partnership for microbicides; LDH, lactate dehydrogenase; LFA, leukocyte function-associated antigen; MIP, macrophage inflammatory protein; MTN, microbicide trials network; MTS, (3-(4,5-dimethylthiazol-2-yl)-5(3-carboxymethoxyphenyl)-2-(4-sulfophenyl)-2H-tetrazolium); MTT, (3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide); N-9, nonoxynol-9; NNRTI, nonnucleoside reverse transcriptase inhibitor; NRTI, nucleoside reverse transcriptase inhibitor; PAMAM, poly(amido amine); PBMC, peripheral blood mononuclear cell; PEG, poly(ethylene glycol); PI, protease inhibitor; PLGA, poly-lactic-co-glycolic acid; PrEP, pre-exposure prophylaxis; RANTES, regulated on activation, normal T cell expressed and secreted protein, also known as CCL5 (Chemokine (C–C motif) ligand 5); SHIV, simian–human immunodeficiency virus; STD, sexually transmitted disease; TC50, toxic concentration 50; TEER, trans epithelial electric resistance; TFV, tenofovir; TNF, tumor necrosis factor; UNAIDS, joint United Nations programme on HIV and AIDS; VOICE, vaginal and oral interventions to control the epidemic; XTT, (2,3-bis-(2-methoxy-4-nitro-5-sulfophenyl)-2H-tetrazolium-5-carboxanilide); 7AAD, 7-aminoactinomycin D. ⇑ Corresponding author at: Laboratorio InmunoBiología Molecular, Hospital General Universitario Gregorio Marañón, CIBER BBN, C/Dr. Esquerdo 46, 28007 Madrid, Spain. Tel.: +34 915868018; fax: +34 915868565. E-mail addresses: [email protected], [email protected] (Ma Ángeles Muñoz-Fernández). 1 Both authors contributed equally to this work. http://dx.doi.org/10.1016/j.antiviral.2014.10.014 0166-3542/Ó 2014 Published by Elsevier B.V.

Please cite this article in press as: Sánchez-Rodríguez, J., et al. Nanotech-derived topical microbicides for HIV prevention: The road to clinical development. Antiviral Res. (2014), http://dx.doi.org/10.1016/j.antiviral.2014.10.014

AVR 3537

No. of Pages 16, Model 5G

11 November 2014 2

J. Sánchez-Rodríguez et al. / Antiviral Research xxx (2014) xxx–xxx

52 53 54 55 56 57 58

5.

4.1.5. Investigation of nanoparticle inhibitory mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1.6. Combination assays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.1.7. Broad spectrum activity of topical microbicides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2. Ex vivo and in vivo evaluation: a proof-of-concept of compound efficacy prior to human clinical trials . . . . . . . . . . . . . . . . . . . . . . . . . . . . Future perspectives, obstacles and conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

00 00 00 00 00 00 00

59 60

61

1. HIV prevention strategies: an urgent necessity

62

As announced in the 2013 UNAIDS report, approximately 35.3 million people are living with HIV worldwide (UNAIDS, 2013). Although during the last few years international and national health programs have achieved a progressive reduction in new HIV infections, 2.5 million new infections and 1.7 million AIDSrelated deaths still occurred worldwide. HIV heterosexual transmission is responsible for approximately 80% of these new infections and is one of the most difficult routes of transmission to control. During transmission, HIV-1 must enter or cross the stratified epithelium of the vagina or the columnar epithelium of the uterine cervix. It is unknown whether infectious virions penetrate mucosal tissues far enough to reach epithelial or subepithelial target cells, which cells are initially infected in addition to CD4+ T cells and to what extent trans-presentation by non-infected dendritic cells contributes to HIV-1 dissemination (Fackler et al., 2014). Women constitute over 50% of the HIV infected population, although in some countries, such as South Africa, this percentage rises to approximately 60% (Carter et al., 2013). The problem is even more alarming in sub-Saharan Africa, where more than six out of every ten infected individuals is female, and young women aged 15–24 years are twice as likely to be infected with HIV compared to men of the same age. Furthermore, HIV/AIDS has become the main cause of death among women aged 15–44 years. In many cases, men and women do not share the responsibility of reducing the risk of HIV transmission because women are excluded from sex-related decision-making, do not receive in-depth sex education

63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88

or have unequal access to preventive methods. That translates into 3000 new infections in women every day. For over 20 years, the scientific community has been working hard to develop new preventive strategies to decrease HIV infection rates, especially in women. To date, despite a great deal of effort, there are no effective vaccines against HIV infection (Munier et al., 2011). Thus, it is urgent and imperative to develop prevention systems capable of stopping the spread of HIV in the most disadvantaged and vulnerable populations, especially during sexual intercourse (Anton, 2012; Stephenson, 2011). It is known that behavioral and structural interventions, such as the use of condoms or circumcision, are effective methods in the prevention of HIV sexual transmission (Padian et al., 2008; Quinn, 2007). Other pre-exposure prophylaxis (PrEP) HIV-preventive strategies, such as the oral administration of antiretroviral (ARV) drugs or topical microbicides, also seem to be feasible and effective approaches against the spread of HIV (Abdool Karim and Baxter, 2014; Balzarini and Van Damme, 2005; Buckheit et al., 2010; Granich et al., 2010). Microbicides are compounds that when applied topically to the vagina and/or rectum can reduce the risk of HIV infection. Microbicides are advantageous over other prevention methods, such as condoms, because they are easier to use and provide users, especially women, with the ability to decide on their utilization. Microbicides are extremely advantageous to highly vulnerable groups, such as sex workers or women who suffer from sexual aggression. Due to their great potential, much effort in the past and present has been put into developing vaginal or rectal microbicides, including gels, rings and other formulations, capable of reducing the risk of HIV sexual transmission (Tables 1–3).

Table 1 Ongoing and planned Phase III and Phase II/IIb clinical trials of microbicide candidates. The only candidate to show efficacy to date is the antiretroviral drug tenofovir formulated in 1% gel form. One trial has shown benefit; another has not. Confirmatory research is ongoing in different trials. A new formulation of tenofovir gel is also being evaluated for rectal used, as a possible HIV prevention tool for anal sex. Other leading candidate in large efficacy trials is dapivirine ring. Currently, there are two ongoing efficacy trials of the dapivirine ring recruiting women from communities across several countries in East and Southern Africa. TDF – Tenofovir disoproxil fumarate; TFV – Tenofovir; FTC – emtricitabine; MSM – men who have sex with men. (Adapted from http://www.avac.org/). Trial name

Phase

Start date

Phase III (safety and effectiveness) FACTS 001 Phase III October 2011

Locations

Population

Candidate(s)

Status/expected completion

South Africa

2900 women

1% TFV gel 4-week vaginal dapivirine ring 4-week vaginal dapivirine ring 1% TFV gel

Ongoing/December 2014 Ongoing/August 2015

IPM 027 (The Ring Study) MTN 020 (ASPIRE)

Phase III April 2012

Rwanda, South Africa, Malawi

1650 women

Phase III July 2012

3476 women

CAPRISA 008

Phase III, October 2012 II

Malawi, South Africa, Uganda, Zimbabwe South Africa

Phase II, IIb (safety, adherence, acceptability, feasibility, effectiveness) MTN 017 Phase II September Peru, South Africa, Thailand, 2013 United States, Puerto Rico MTN 019 MTN 023/IPM 030

Phase II Phase II

July 2011 March 2014

MTN 024/IPM 031

Phase II

FACTS 002

Phase II

December 2013 To be determined

United States United States South Africa

700 women

186 transgender, MSM

Daily oral TDF/FTC, Reformulated rectal 1% TFV gel 1% TFV gel Dapivirine vaginal ring

Ongoing/December 2014 Ongoing/February 2015 Ongoing/June 2016

384 women HIV-negative adolescent females 96 postmenopausal women Dapivirine vaginal ring

Ongoing/August, 2015 Ongoing

60 young women (16– 17 years)

Ongoing

Vaginal TFV 1% gel

Ongoing/May 2014

Please cite this article in press as: Sánchez-Rodríguez, J., et al. Nanotech-derived topical microbicides for HIV prevention: The road to clinical development. Antiviral Res. (2014), http://dx.doi.org/10.1016/j.antiviral.2014.10.014

89 90 91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 117

AVR 3537

No. of Pages 16, Model 5G

11 November 2014 3

J. Sánchez-Rodríguez et al. / Antiviral Research xxx (2014) xxx–xxx

Table 2 Current Phase I microbicide clinical trials. The majority of microbicide candidates currently in clinical trials are formulated with ARV drugs and most of them have been gels. However, there are several other candidates in early phases of development. In addition, it may be useful to develop non-ARV based candidates and candidates delivered in other ways (e.g., a film or fast dissolving tablets). Trial name

Phase

Start date

Locations

Population

Phase I (safety, adherence, acceptability, feasibility) CONRAD A10-114 Phase I March 2012 United States, Dominican Republic CONRAD 122 Phase I To be United States determined CONRAD 118 Phase I December United States 2013 CONRAD 117 Phase I February United States 2013 CONRAD 114 Phase I March 2012 Dominican Republic, United States Project Gel Phase I March 2014 United States, Puerto Rico

NNRTI Microbicide Phase I Gel MTN-014 Phase I IPM-034 Phase I Population Council Phase I #558 MTN-011 Phase I

Candidate(s)

Status/expected completion

March 2014

United States

April 2014 November 2013 March 2014

United States Belgium

Ongoing/January 2014 40 men and women SILCS diaphragm and 1% TFV gel Protocol in development 54 women, HIV negative 1% TFV gel, Contraceptive Vaginal Ring, Ongoing/November Metronidazole, Terconazole 2014 48 women, HIV negative TFV/FTC and TFV only fast dissolve Ongoing/January vaginal tablets 2014 72 women 1% TFV gel, Depo-provera, Oral Ongoing contraceptive 240 MSM 1% TFV rectal gel once daily for 7 days Ongoing/March 2014 and placebo rectal gel prior to rectal intercourse 30 women, HIV negative PC-1005 gel, HEC gel Ongoing/March 2015 28 women 1% tenofovir reduced-glycerin gel Ongoing 40 women Dapivirine Ring Ongoing/May 2014

United States

women

PC-1005 gel

United States

80 women, men, HIV negative women

1% TFV gel

90 women

TFV and TFV/LNG vaginal ring

CONRAD 123

Phase I

November 2012 2014

CONRAD 128

Phase I

Q4 2013

72 women, HIV negative 1% TFV gel, DMPA

Ongoing/March 2015 Ongoing/April 2015

TFV gel/SILCS Diaphragm

Protocol in development Pending

Table 3 Open label and other clinical trials of microbicide candidates. Various other clinical studies and sub-studies are currently ongoing to better understand, for instance, the adherence and use of the study products or the sexual behavior of users during their participation in the trials. Trial name

Open label MTN 025

Phase

Start date

Locations

Population

Vaginal dapivirine ring

Planned

N/A

Ongoing/ May 2013

MTN 016 (EMBRANCE) IPM-33

Open label

August 2013

IPM-29

Open label

February 2013

United States

Pre-clinical

Q4 2013

To be determined Eastern Virginia Medical School/To be determined

TFV

Planned 2014

South Africa

N/A

Primary analysis completed Primary analysis completed Ongoing/ March 2014

MTN 003B

MTN 015

MTN 003 (VOICE) July 2010 community sub-study MTN 003 (VOICE) bone mineral density sub-study Seroconverter protocol

November Uganda, Zimbabwe 2009 August 2008

140 women/men

275 Participants (includes women in VOICE, their male partners, CAB members and community stakeholders) 518 women

Malawi, South Africa, 500 HIV-positive women Uganda, Zambia, Zimbabwe

118

2. History of the microbicide pipeline

119

Many agents with antiviral activity have been evaluated as microbicidal candidates. The first tested microbicides were

120

Status/ expected completion

Open label/phase IIIb ASPIRE follow on Open label/pregnancy October exposure registry 2009

Other Effects of Tenofovir on CAPRISA samples MTN 003C

Malawi, South Africa, Uganda, Zimbabwe South Africa, Uganda, 800 women, men, HIV negative United States, Zimbabwe United States 160 women/men

Candidate(s)

Ongoing No ring, condom only/ placebo vaginal ring + female condom used Silicone elastomer vaginal Ongoing/ ring October 2013

Oral TDF; oral TDF/FTC

N/A

surfactants, which are compounds that act as detergents and are able to destroy the integrity of lipid membranes. The most well known is the non-specific microbicide nonoxynol-9 (N-9) (Roddy et al., 1998; Van Damme et al., 2002). This compound was able

Please cite this article in press as: Sánchez-Rodríguez, J., et al. Nanotech-derived topical microbicides for HIV prevention: The road to clinical development. Antiviral Res. (2014), http://dx.doi.org/10.1016/j.antiviral.2014.10.014

121 122 123 124

AVR 3537

No. of Pages 16, Model 5G

11 November 2014 4 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 186 187 188 189 190

J. Sánchez-Rodríguez et al. / Antiviral Research xxx (2014) xxx–xxx

to destroy the virus by damaging the viral envelope (Bourinbaiar and Fruhstorfer, 1996). However, clinical trials showed that N-9 increased the risk of HIV infection (Roddy et al., 1998). Another surfactant, C31G (SAVVY), which was shown to be safe in vitro and exhibited broad-spectrum activity against bacteria and herpes simplex virus-2 (HSV-2), was analyzed against HIV in three independent clinical trials in Ghana and Nigeria (Ballagh et al., 2002; Bax et al., 2002; Feldblum et al., 2008; Krebs et al., 1999; Mauck et al., 2004; Thompson et al., 1996). These trials were suspended because the rates of infection were higher in the surfactant-treated group compared to the placebo group. Currently, the surfactant candidate sodium lauryl sulfate (Invisible Condom; Université Laval, Quebec, Canada) is being investigated (Bestman-Smith et al., 2001; Mbopi-Keou et al., 2010; Piret et al., 2002). It works as an ‘invisible condom’ that could cover the vaginal wall as a liquid at room temperature and subsequently transform into a gel at body temperature, inhibiting the transmission of HIV-1 and other sexual transmitted diseases (STD). The results from a randomized, double blind, placebo-controlled Phase II study in Cameroonian women revealed that sodium lauryl sulfate was well tolerated and acceptable (Mbopi-Keou et al., 2010). Further phases of clinical development are planned (Gupta, 2013). Together with surfactants, many polyanionic compounds have been analyzed as microbicides (Baranova et al., 2011; Pirrone et al., 2011). The anti-HIV activity of these compounds is associated with the establishment of electrostatic interactions between the viral envelope proteins and the anionic functional groups of the compounds. This interaction prevents the binding of the virus to the target cell, blocking viral entry. Three of these polyanionic compounds, Carrageenan (Carraguard), cellulose sulfate and PRO2000, were effective in vitro and in preclinical trials, although none showed any efficacy in clinical trials in humans (Gibson and Arts, 2012; Huskens et al., 2009; Keller et al., 2010; Pirrone et al., 2011; Reina et al., 2010). Surprisingly, in some cases, such as with cellulose sulfate, higher rates of HIV infection were found in the treated group compared to the placebo group (Tao et al., 2008). Several rationales were proposed to explain the failure of these polyanionic compounds, including their lower inhibitory effects in comparison to other antivirals, poor absorption by the mucosa, reduced efficacy in the presence of seminal plasma, or the induction of local inflammation and changes in vaginal microflora, all of which could facilitate HIV-1 infection (Pirrone et al., 2011). Currently, the leading candidates in the microbicide pipeline are ARV (Chirenje et al., 2010; Granich et al., 2010; Klasse et al., 2008; Omar and Bergeron, 2011; Shattock and Rosenberg, 2012). These compounds exhibit potent and specific anti-HIV activity and are safe, as demonstrated by their continuous use in highly active antiretroviral therapy (HAART). Currently, more than thirty ARV targeting different steps of the viral cycle have been approved by the U.S. Food and Drug Administration (FDA) for use in HAART (Fig. 1). HIV inhibitors, including entry inhibitors, nucleoside and non-nucleoside reverse transcriptase inhibitors (NRTI and NNRTI, respectively) or integrase and protease inhibitors have been used as potential microbicide candidates. The only candidate to show high efficacy to date is the NRTI Tenofovir (TFV) (Abdool Karim et al., 2010; Gengiah et al., 2012; Mertenskoetter and Kaptur, 2011). TFV formulated as a 1% gel was tested in the CAPRISA 004 trial in 889 South African women, resulting in a 39% overall reduction in the women’s risk of contracting HIV infection via vaginal sex (Network, 2010). However, other clinical trials failed, such as VOICE, a five-armed proof-of-concept clinical trial that enrolled more than 5000 women in South Africa, Uganda and Zimbabwe (Friend and Kiser, 2013). The VOICE study was prematurely stopped in 2011 due to lack of efficacy. It concluded that none of the interventions tested (daily oral TFV, daily oral TFV/emtricita-

bine or daily TFV gel) reduced the risk of HIV transmission. Some reasons have been provided to explain the apparent lack of efficacy, particularly the low adherence. Most participants did not use the tested interventions daily as recommended. Thus, confirmatory research is ongoing, and trials with other formulations of TFV are planned (Network, 2013a). For instance, CAPRISA 008 will assess the feasibility and effectiveness of providing the TFV gel in a clinic setting. A new reduced glycerin formulation of the TFV gel is being evaluated for rectal use in the MTN 017 trial as an HIV prevention system for rectal transmission (Network, 2013b). This is particularly relevant to homosexual communities and other men who have sex with men, especially when the risk associated with HIV transmission through unprotected receptive anal intercourse is 1.7% per act (versus 0.08% in case of unprotected vaginal intercourse). This is due to anatomical differences between vaginal and rectal tissues. The rectal epithelium is formed by a single layer of cells as opposed to the multilayer squamous epithelium found in the vagina and ectocervix. In addition, the gastrointestinal tract is populated with a large number of HIV-1 infectable cell subsets (van Marle et al., 2007). All these factors make the rectal route a more susceptible route for HIV infection. Several rectal microbicide clinical trials testing others antiretrovirals such as UC781 have been conducted, as deeply reviewed by Rohan et al. (2013). Another candidate in large-scale efficacy trials is the NNRTI dapivirine. It was formulated as a vaginal ring, and currently there are two ongoing phase III clinical trials (ASPIRE and IPM 027) in Africa (Microbicides, 2012; Network, 2012). Another phase I clinical trial, MTN 013/IPM 026, has studied the effectiveness of the combination of dapivirine and maraviroc as a microbicidal vaginal ring (Fetherston et al., 2012). Results were presented this year at the 21st Conference on Retroviruses and Opportunistic Infections (CROI). MTN-013/IPM 026 found that the ring was safe in women. In addition, dapivirine was able to block HIV infection, though levels of maraviroc were not sufficient to have a similar effect (Chen et al., 2014, (CROI)). Because CCR5-tropic viral strains are dominant during HIV-1 sexual transmission, entry inhibitors, such as CCR5 antagonists, have also been evaluated as microbicide candidates (Maeda et al., 2012). CCR5 antagonist CMPD167, a cyclopentane-based compound used as a vaginal gel that has provided protection from vaginal SHIV challenge in macaques, and maraviroc, a small molecule that binds to the CCR5 co-receptor and blocks HIV-1 entry into cells, are being tested in humans. A proof-of-concept study in macaques showed that maraviroc could play a role in protecting women from HIV-1 infection if developed as a vaginal microbicide (Veazey et al., 2010). A major challenge in considering the use of CCR5 inhibitors as topical microbicides is their inability to block the entry of X4-tropic viruses. Although infection via the CXCR4 co-receptor is less important in sexual transmission, X4-tropic viruses are present in vaginal fluids and semen during acute HIV infection (Grivel et al., 2011). Thus, the development of broad spectrum topical microbicides capable of preventing HIV transmission regardless of viral tropism is crucial. There are several other novel ARV-based and non-ARV-based candidates in the early phases of preclinical development. For instance, the cyclic antimicrobial peptide Retrocyclin RC-101 (Gupta et al., 2013), the NRTI saquinavir (Stefanidou et al., 2012) and rilpivirine (De Clercq, 2012), the fusion inhibitor cyanovirin (Tsai et al., 2003) and zinc-derived compounds (Kenney et al., 2011) are also under development. Interestingly, griffithsin, a carbohydrate-binding agent (CBA) extracted from red algae, is expected to enter in a phase I safety and acceptability trial in the next year (MTN-038) (Emau et al., 2007; Hillier, 2013). A summary of the clinical trials of microbicide candidates is shown in Tables 1– 3. The outcomes of these trials currently underway will significantly impact on the development and search of new candidates.

Please cite this article in press as: Sánchez-Rodríguez, J., et al. Nanotech-derived topical microbicides for HIV prevention: The road to clinical development. Antiviral Res. (2014), http://dx.doi.org/10.1016/j.antiviral.2014.10.014

191 192 193 194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256

AVR 3537

No. of Pages 16, Model 5G

11 November 2014 J. Sánchez-Rodríguez et al. / Antiviral Research xxx (2014) xxx–xxx

5

Fig. 1. Therapeutic targets of the HIV life cycle. Viral entry requires the binding of gp120 to the cellular receptor CD4 and a co-receptor (CCR5 or CXCR4). Co-receptor binding causes changes in gp120 and the gp41 transmembrane peptide, leading to membrane fusion. CCR5 inhibitors, such as maraviroc (CelsentriÒ), prevent viral binding to this co-receptor. Fusion inhibitors as Enfuvirtide (FuzeonÒ) bind to gp41 and prevent the conformational changes required for viral cell fusion. In the cytoplasmic compartment, the internalized viral core and viral enzyme reverse transcriptase copies the viral RNA into a DNA strand which is transported into the nucleus. Analogue reverse transcriptase inhibitors and non-nucleoside inhibitors (NRTI and NNRTI), such as tenofovir (VireadÒ), abacavir (ZiagenÒ) and efavirenz (SustivaÒ), inhibit this process. In the nucleus, the viral integrase incorporates the viral DNA into the cellular genome, a step blocked by integrase inhibitors, such as raltegravir (IsentressÒ). The integrated proviral DNA is transcribed to generate complete viral RNA sequences, and multiple mRNAs are translated into proteins using cellular machinery. The viral RNA and viral proteins are transported to the cell membrane, where they are assembled with cellular factors within the viral particle. Budding of viral particles occurs simultaneously with the cleavage of precursor Gag and Gag-Pol by the viral protease, resulting in a mature virion. Protease inhibitors, such as atazanavir (ReyatazÒ) and darunavir (PrezistaÒ), block the budding step and prevent the formation of mature virions.

257 258 259 260

However, there are some limitations that could explain the modest success of the use of antiretroviral drugs in preventive approaches against HIV, such as their extreme hydrophilic and hydrophobic nature or their low enzymatic and chemical stability.

261

3. Nanotechnology: a new field in the design of microbicides

262

In recent years, nanotechnology has emerged as a new research area capable of providing solutions to the limitations found in the development of preventive strategies against HIV. This field is focused on designing, engineering and fabricating new materials at the nanosize level. According to the National Nanotechnology Initiative, nanotechnology involves study of materials/architectures of size 1–100 nm in at least one dimension. Nevertheless, materials with size up to several hundred nanometers are also included under nanotechnology. These materials can be classified

263 264 265 266 267 268 269 270

into several groups, such as liposomes, polymers, dendrimers, nanocrystals, nanofibers and metal nanoparticles, according to the material used in their assembly (Table 4). In biomedical applications, nanotechnology has enabled the development of novel nanodrugs and compounds with useful pharmacological advantages in the therapy and diagnosis of cancer and infectious diseases, including HIV (Kim et al., 2010a; Kim and Read, 2010; Siccardi et al., 2013). Various nanosystems have been used in different strategies for the treatment and prevention of HIV-1 infection (Date and Destache, 2013; Mahajan et al., 2012; Mallipeddi and Rohan, 2010; Mamo et al., 2010). Some nanoplatforms, due to their nanometric size, are versatile drug delivery systems because of their capacity to overcome corporal barriers and deliver drugs to specific cells or intracellular compartments. Different methods of cargo delivery include encapsulation in the particle core or absorption onto their functionalized surface, thereby improving the pharma-

Please cite this article in press as: Sánchez-Rodríguez, J., et al. Nanotech-derived topical microbicides for HIV prevention: The road to clinical development. Antiviral Res. (2014), http://dx.doi.org/10.1016/j.antiviral.2014.10.014

271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287

AVR 3537

No. of Pages 16, Model 5G

11 November 2014 6 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353

J. Sánchez-Rodríguez et al. / Antiviral Research xxx (2014) xxx–xxx

cological properties of several ARV. For instance, nanosuspensions of rilpivirine stabilized by polyethylene–polypropylene glycol and PEGylated tocopheryl succinate ester resulted in extended sustained release of the compound in animal models, suggesting that nanoscale drug delivery may potentially reduce dosing frequency and improve adherence (Baert et al., 2009). Nanoplatforms can also be used to control the release of nucleic acids in gene therapy against HIV, or as new antiviral agents due to their intrinsic antiviral properties (Amiji et al., 2006; Briz et al., 2012; Cordoba et al., 2013b,c; Chonco et al., 2012; das Neves et al., 2010; de Las Cuevas et al., 2012; Farokhzad, 2008; Jimenez et al., 2010; Mallipeddi and Rohan, 2010; Perise-Barrios et al., 2014; Sanchez-Nieves et al., 2014; Weber et al., 2008, 2012; Meng et al., 2011; Steinbach et al., 2012). These nano-architectures have also been used in different approaches to design novel topical microbicides (Date and Destache, 2013; Ensign et al., 2014). Polymeric biodegradable nanoparticles, liposomes, nanofibers and inorganic nanoplatforms are being explored as platforms for vaginal drug delivery. TFV and tenofovir disoproxil fumarate nanoparticles composed of a blend of poly-lactic-co-glycolic acid (PLGA) and pH-responsive polymethylmethacrylate (Eudragit) polymer, dapivirine-loaded poly(epsilon-caprolactone) nanoparticles, raltegravir-efavirenzloaded PLGA nanoparticles or PLGA nanoparticles with encapsulated PSC-RANTES (an antagonist of the CCR5 co-receptor) have already been tested in animal models against HIV (Belletti et al., 2012; Chaowanachan et al., 2013; das Neves et al., 2012a, 2013; Ham et al., 2009; Zhang et al., 2011; Ensign et al., 2014). PSC-RANTES encapsulation within PLGA nanoparticles provided sustained release and increased uptake into ex vivo cervical tissue compared to soluble PSC-RANTES. Other major study has shown that PSCRANTES is effective and provides potent protection against vaginal challenge in rhesus macaques, indicating its potential as microbicide candidate (Lederman et al., 2004). Liposomes composed of phospholipid bilayer structures were investigated as antiviral agents due to discovery that certain lipid compositions can bind to HIV particles and modulate their infectivity. Moreover, lipid-based nanoparticles have been designed for vaginal delivery of drugs or small interfering RNA to silence HIV gene expression and prevent the establishment of primary infection (Malavia et al., 2011; Wang et al., 2012). The ability of nanofibers, which are spaghetti-like compounds made from materials, such as polysaccharides, peptides or polymers with diameters ranging from 1 to 1000 nm, to deliver different microbicides for HIV prevention has been explored. Maravirocloaded nanofibers, nanofibers made of cellulose acetate phthalate (CAP) that acts as an HIV entry inhibitor and CAP fibers loaded with TFV have shown great potential in inhibiting HIV infection (Gunn and Zhang, 2010; Huang et al., 2012). Furthermore, the antiviral properties of inorganic metals, such as silver, gold and zinc, have been exploited in the search for novel topical microbicides. Polyvinyl pyrrolidone-coated silver nanoparticles inhibited viral fusion and demonstrated antiviral activity against various HIV strains, including drug-resistant strains (Lara et al., 2010a,b). Gold nanoparticles contain free thiol groups on their periphery that permit their easy functionalization with various biomolecules. Multivalent oligomannosides-coated gold nanoparticles and anionic sulfate-functionalized gold nanosystems have demonstrated inhibitory activity against HIV in vitro (Di Gianvincenzo et al., 2010; Vijayakumar and Ganesan, 2012). Dendrimers are the most advanced group of nanotech-derived microbicides (Lazniewska et al., 2012; Parboosing et al., 2012; Svenson and Tomalia, 2005). Dendrimers are a class of hyperbranched polymeric nano-structures that are much smaller than traditional nanoparticles, with narrow polydispersity and precise numbers of terminal groups. Dendrimer structure and composi-

tion, including the number of branching units and the composition and number of surface functional groups, can be engineered to obtain compounds with the desired physicochemical and biological properties. The peripheral functional groups can be designed to directly interact with target cells or pathogens, conjugate drugs, or complex with other biomolecules, such as peptides or siRNA. One dendrimer based-compound, VivaGel, is the only nanotechderived microbicide candidate that has reached human clinical trials (Bernstein et al., 2003; Gong et al., 2005; Jiang et al., 2005; Telwatte et al., 2011; Tyssen et al., 2010). The active ingredient in VivaGel is the SPL7013 dendrimer, which is a polylysine anionic dendrimeric agent. It has demonstrated great potency against both HIV-1 and HSV-2 (Price et al., 2011; O’Loughlin et al., 2010), although the phase I clinical trial evaluating its antiviral activity and microbiota tolerance revealed evidence of mild irritation after repeated vaginal use (McGowan et al., 2011). VivaGel is not currently being tested in clinical trials as a topical microbicide against HIV; nevertheless, it is in phase II trials as a treatment for bacterial vaginosis. Other dendrimeric compounds are in development for use as vaginal microbicides (Date and Destache, 2013; Navath et al., 2011). Dendrimers containing a carbosilane core and different anionic ended functional groups have been studied as anti-HIV compounds (Chonco et al., 2012; Galan et al., 2014; Garcia-Gallego et al., 2012; Rasines et al., 2012; Vacas Cordoba et al., 2013). These dendrimers have shown high biosafety in human epithelial cell lines derived from the female genital tract and in primary human blood cells (PBMC), blocked the entry of both X4- and R5-tropic HIV-1 isolates into epithelial cells while protecting the epithelial monolayer from cell disruption, reduced HIV-1 infection of PBMC and were safe in mouse and rabbit models, with no promotion of irritation or vaginal lesions (Chonco et al., 2012; Vacas Cordoba et al., 2013). Anionic dendrimers comprising benzhydryl amide cores and lysine branches and functionalized with naphthalene disulfonic acid or 3,5-disulfobenzoic acid groups have demonstrated activity against HIV-1 and HSV-2 infection, preventing viral entry and HIV envelope-mediated cell-to-cell fusion (Telwatte et al., 2011; Tyssen et al., 2010). Polycationic ‘‘viologen’’ based dendrimers have also shown anti-HIV activity by interaction with the CXCR4 co-receptor (Asaftei et al., 2012). Other dendrimers, such as four generation-PAMAM branched-naphthalene disulfonic surface groups ending with SPL2923 or polysulfonated dendrimer BRI2923, have shown inhibitory effects against HIV-1 not only at the entry step but also at later stages of the viral replicative cycle (Witvrouw et al., 2000). Multivalent dendrimeric compounds coated with carbohydrates expressed on immune cells inhibited HIV-1 infection by targeting viral entry, whereas multivalent-glycodendrimeric compounds prevented HIV transmission via DCSIGN on dendritic cells (Garcia-Vallejo et al., 2013; Rosa Borges et al., 2010). Although nanoparticles are under development as useful and powerful tools in the field of antiviral microbicides, they display a series of limitations that should be addressed because they can affect the transfer of these platforms to clinical trials. These limitations include toxicity, the appearance of non-desirable biological interactions, bioaccumulation, the enzymatic degradation of these nanosystems, their penetrance and absorption features into different tissues and their high manufacturing cost; this last limitation is especially a problem for scale-up production (Mamo et al., 2010).

354

4. Moving towards pre-clinical development of nanotechderived microbicides: the main stages

412

Rigorous pre-clinical evaluation of vaginal or rectal candidate microbicides is essential for the selection of the best compounds for continued development. This involves a series of basic steps

414

Please cite this article in press as: Sánchez-Rodríguez, J., et al. Nanotech-derived topical microbicides for HIV prevention: The road to clinical development. Antiviral Res. (2014), http://dx.doi.org/10.1016/j.antiviral.2014.10.014

355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411

413

415 416

Architecture

Features

AVR 3537

11 November 2014

Nanocompound

Candidates explored in vivo  Anti CCR5 siRNA-loaded liposome (Kim et al., 2010a)  Cardiolipin liposome (Malavia et al., 2011)

Liposomes

 Biocompatible, completely biodegradable and non-immunogenic  Encapsulation of hydrophobic, amphipathic and hydrophilic drugs  Ability to protect encapsulated drugs from the external environment and to act as sustained release depots  Available technologies for large-scale production

Dendrimers

   

Controlled synthesis and narrow polydispersity Different molecules/drugs could be conjugated to a single dendrimer Due to their hyperbranched architecture, they can interact in a multivalent way They can be tailored by manipulating their structure or composition and number of surface functional groups to obtain compounds with desired properties

 Vivagel (SPL7013 dendrimer) (O’Loughlin et al., 2010)  Sulfonated carbosilane dendrimer 2G-S16 (Chonco et al., 2012)  Tenofovir-emtricitabine-loaded thiopyridine terminated G4PAMAM dendrimer (Navath et al., 2011)

Polymeric nanoparticles

   

There are a wide range of biodegradable and non-biodegradable materials for desired application Sustained release of encapsulated agents Improved pharmacokinetic, solubility and stability of encapsulated agents They can be used in targeted-delivery by tailoring their surface

Nanofibers

   

A wide variety of materials can be assembled into nanofibers Large surface area and many active surface sites Delivery and sustained drug release Easy production process: electrospinning

 PSC-RANTES loaded PLGA-nanoparticles (Ham et al., 2009)  siRNA-loaded PLGA nanoparticles (Steinbach et al., 2012)  Tenofovir-loaded pH-sensitive PLGA nanoparticles (Zhang et al., 2011)  Tenofovir-loaded chitosan nanoparticles (Meng et al., 2011)  Tenofovir loaded hybrid chitosan/PLGA nanoparticles (Belletti et al., 2012)  Dapivirine-loaded polycaprolactone nanoparticles (das Neves et al., 2012a,b)  Tenofovir-loaded cellulose acetate phthalate nanofibers (Huang et al., 2012)

Inorganic nanoparticles

   

High surface-to-volume ratios Manipulable optical and surface characteristics Multiple ligands/drugs could be conjugated to surface using appropriate chemistry Applications in electrooptic devices, biomedical imaging, diagnostic or therapeutic field

Nanocrystal

    

Drug molecules are micronized in nano-scale crystals using top-down/bottom-up techniques Nanocrystals are composed of 100% drug; there is no carrier material Increase of dissolution velocity by surface area enlargement Increase in saturation solubility and bioavailability There are feasible technologies for large-scale production

J. Sánchez-Rodríguez et al. / Antiviral Research xxx (2014) xxx–xxx

 Silver nanoparticles coated condoms (Mohammed Fayaz et al., 2012)

No. of Pages 16, Model 5G

7

Please cite this article in press as: Sánchez-Rodríguez, J., et al. Nanotech-derived topical microbicides for HIV prevention: The road to clinical development. Antiviral Res. (2014), http://dx.doi.org/10.1016/j.antiviral.2014.10.014

Table 4 Important features and schematic structure of nanosystems used in the development of nano-based HIV topical microbicides. Different nano-architectures (liposomes, dendrimers, polymeric nanoparticles, nanocrystals, nanofibers and inorganic nanoparticles) are being explored as novel microbicide candidates. The most advanced in the microbicide pipeline are liposomes, polymers and dendrimers. Liposomes are spherical vesicles that comprise one or more lipid bilayer structures enclosing an aqueous core. They are high-potency carriers that protect encapsulated drugs from degradation. Polymeric nanoparticles are basically solid polymers with drugs embedded in the polymer matrix or with an inner space loaded with the drug of interest. Dendrimers are highly branched polymers with a controlled three-dimensional structure around a central core that can be easily functionalized. (Adapted from Date and Destache (2013), Biomaterials).

AVR 3537

No. of Pages 16, Model 5G

11 November 2014 8 417 418 419

J. Sánchez-Rodríguez et al. / Antiviral Research xxx (2014) xxx–xxx

designed to evaluate and guarantee safety and efficacy (Buckheit and Buckheit, 2012). There are different parameters that need to be studied in several in vitro and in vivo models (Fig. 2).

420

4.1. Identification of novel candidates: in vitro screening

421

4.1.1. Characteristics of ideal nanotech-derived microbicide candidates The discovery and development of new nanocompounds as microbicides with antiviral properties is difficult because these agents need to exhibit a series of desirable properties. They must be non-toxic, safe for prolonged use and highly effective against HIV-1 and other STDs (Fig. 3). They must also be easy to use, odorless and colorless, inexpensive and easy to manufacture and scaleup. Moreover, they must minimally damage the mucosa, be safe to the microflora found at the target mucosal tissue and not interfere with the efficacy and safety of other therapeutic agents, such as contraceptives. Because microbicides should diminish viral transmission both vaginally and rectally, different formulations need to be designed to prevent HIV transmission by these routes (Morrow and Hendrix, 2010; Turpin, 2011). The microbicide’s effectiveness should not be influenced by factors, such as seminal plasma or pH changes, that occur in the vagina during sexual intercourse. Thus, rational design of these nanoplatforms could be crucial for optimal performance in the vaginal and rectal environments (Fahmy et al., 2008). Properties such as mucus-

422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439

penetration ability and sustained drug release capacity are important for improving vaginal drug delivery in HIV prophylactic approaches (Ensign et al., 2014). For instance, PEG surface density has been shown to be important for efficient mucus penetration (Cu and Saltzman, 2009; Tang et al., 2009). Recent progress in the design of biodegradable compounds, together with the regulation of nanosystem design and preparation, makes it possible to precisely manipulate the nanoparticle’s architecture, molecular weight and chemical composition, resulting in predictable tuning of their biocompatibility as well as their delivery.

440

4.1.2. Antiviral assays Initial screening assays provide assessments of the compound’s efficacy (EC50: 50% reduction in HIV production from treated and untreated cells) and allow the identification of inhibitory compounds with direct antiviral activity or with the capacity to inhibit a specific step of the HIV life cycle. Human epithelial cell lines from appropriate target tissues, such as the vaginal mucosa-derived VK2/E6E7, uterine mucosa-derived HEC-1A or colorectal epithelium Caco-2 cell lines are commonly used in the study of transepithelial infection process (Gali et al., 2010a,b). Although infection of epithelial cells has been previously described, it is most probably not productive (Dezzutti et al., 2001). True early target cells of sexually acquired HIV include intra-epithelial cells such as Langerhans cells, subepithelial macrophages or CD4+ T lymphocytes. These

450

Fig. 2. Algorithm for selection and evaluation of HIV microbicide candidates. It comprises the main steps for a rationale preclinical development of novel nano-based anti-HIV agents, including the evaluation of compound efficacy and toxicity in relevant cell-based systems, the range and mechanism of action of candidate compounds, the interactions of the compound in combination with other approved antiretrovirals, in vitro and in vivo pharmacokinetic, pharmacodynamics and safety evaluations and efficacy studies in sensitive models such as non-human primate or humanized mice. Adapted from Science Working Group of the Microbicide Initiative, The Science of Microbicides: Accelerating Development, 2002.

Please cite this article in press as: Sánchez-Rodríguez, J., et al. Nanotech-derived topical microbicides for HIV prevention: The road to clinical development. Antiviral Res. (2014), http://dx.doi.org/10.1016/j.antiviral.2014.10.014

441 442 443 444 445 446 447 448 449

451 452 453 454 455 456 457 458 459 460 461 462 463

AVR 3537

No. of Pages 16, Model 5G

11 November 2014 J. Sánchez-Rodríguez et al. / Antiviral Research xxx (2014) xxx–xxx

9

Fig. 3. Potential activity of topical microbicides in vaginal epithelium. Microbicides should have not only anti-HIV activity, but they must be effective against other STDs that can increase the risk of HIV transmission by increasing the risk of epithelial inflammation and ulceration. Polyanionic compounds can disrupt the viral envelop or block cellvirus fusion. An acidic pH is necessary for natural vaginal protection against pathogens. Therefore it is important that microbicides do not alter the normal vaginal microflora. When the virus crosses the epithelial barrier it is taken up dendritic cells prior to infection of T cells and the recruitment of additional susceptible cells (adapted from Shattock et al., Nat Rev Microbiol, 2003).

464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489

490 491 492 493 494 495 496 497 498 499 500

PBMCs subets and the CCR5-expressing TZM-bl reporter cell line are the best choice in the primary evaluation of microbicide candidates. Because R5-tropic viral isolates are mainly involved in HIV sexual transmission, the evaluation of the antiviral activity of nanoparticles against these viral strains is relevant because the majority of sexually-transmitted infections are via the CCR5 coreceptor. CD4+ T cells can be found along the lamina propia in the human vagina (endo- and ectocervix), often under the basal membrane (Edwards and Morris, 1985). Most are memory T cells that have increased CCR5 expression compared to T cells circulating in the peripheral blood (Hladik and McElrath, 2008; Hladik et al., 2007; Zhang et al., 1998). Nevertheless, it is important to remember that CXCR4-tropic viral strains are also present in vaginal fluids and semen. Therefore, to assure broad spectrum and potent antiviral activity, microbicide candidates should be tested against CCR5-, CXCR4- and dual-tropic viral strains. The most common R5-tropic laboratory strains are HIV-1JR-CSF, HIV-1BaL and HIV1NL(AD8) (B subtype) and the most common dual-tropic strain is HIV-1CBL23. All of these strains infect human primary cells and cell lines productively. However, to evaluate the efficacy of nanoparticles, testing their antiviral activity against different subtypes of primary clinically isolated viral strains is also necessary. Owing to the difficultly involved in using primary viral strains, working with established laboratory viral strains is an easier and quicker first step of in vitro primary screening of novel candidates.

4.1.3. Safety and biocompatibility of nanoparticles Once the inhibitory potency of new compounds is established, it is important to evaluate their safety and 50% toxic concentration (TC50) in different primary cells and cell lines. Different commercial assays are available for assessing nanoplatform toxicity in cells (Duncan and Izzo, 2005), the most common being the MTT assay. Other closely related colorimetric assays, including XTT and MTS, are also used. Studying cell membrane integrity is another common method to measure cell viability and cytotoxic effects. Vital dyes are frequently employed, such as trypan blue, 7-aminoactinomycin D (7-AAD) or propidium iodide. Alternatively, enzymatic

assays can be used to monitor the passage of molecules though the membrane, such as the LDH assay. Nanoparticle cytotoxicity depends on different features, such as nanoparticle size, surface charge and structure hydrophobicity/ hydrophilicity as reviewed by Date and Destache (2013). For instance, in cationic nano-architectures, the chemical structure of used cationic lipids can dramatically impact their efficacy, cellular uptake and toxicity. Nanosystems composed of cationic lipids with single alkyl chain were more toxic and incapable of transport through cervicovaginal mucus in comparison to nano-architectures composed of cationic lipids with two alkyl chains (das Neves et al., 2012b). Molecular weight is another important property that governs performance and toxicity of nano-architectures. In case of polyethylenimines (PEI), polymer molecular weight has an enormous influence on their transfection capability, antiviral activity and cytotoxicity. Size dependent toxicity has been also observed for silver nanoparticles. 15 nm sized-silver nanoparticles were significantly more toxic to macrophages than 30–55 nm silver nanoparticles (Carlson et al., 2008). On the contrary, PLGA nanoparticles did not show size dependent toxicity in macrophages, showing that, other factors, such as material type, are also crucial in size dependent responses (Xiong et al., 2013). Surface characteristics such as surface charge, hydrophilicity or surface functional groups have significant impact on nanoparticles behavior. Several studies have shown that positively charged nanocarriers have higher cytotoxicity as compared to anionic and neutral nanocarriers. Oral administration of positively charged dendrimers to mice was found to cause adverse effects such as hemobilia and splenomegaly, whereas negatively charged dendrimers did not show adverse effects and were well tolerated even at 10-fold higher dose (Thiagarajan et al., 2013). The decoration of nanosystem surfaces with biocompatible groups and moieties, such as sugars or as PEG, could also improve nanocompound biocompatibility. Therefore, a thorough understanding of the relationships between the physicochemical properties and the behavior of nanomaterials in biological systems is highly necessary for designing safe and efficacious nano-based microbicides. Novel techniques such as quantitative structure–

Please cite this article in press as: Sánchez-Rodríguez, J., et al. Nanotech-derived topical microbicides for HIV prevention: The road to clinical development. Antiviral Res. (2014), http://dx.doi.org/10.1016/j.antiviral.2014.10.014

501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539

AVR 3537

No. of Pages 16, Model 5G

11 November 2014 10 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603

J. Sánchez-Rodríguez et al. / Antiviral Research xxx (2014) xxx–xxx

activity relationship (QSAR) methods can help to establish such relationships (Burello and Worth, 2011). Another important point to consider when developing a vaginal microbicide is that the compound should not alter the normal vaginal flora or have spermicidal activity. Normal vaginal flora is composed of different Lactobacillus species, mainly Lactobacillus crispatus and Lactobacillus iners but also Lactobacillus jensenii and Lactobacillus , as well as other minority populations such as Atopobium vaginae, Megasphaera or Leptotrichia species (Martin, 2012). Finally, it is important that the compounds remain stable and maintain their antiviral activity in an acidic medium, such as that present in the human female vagina (pH  4–5.8). Factors, such as seminal plasma or pH changes, that occur in the vagina during the sexual act can diminish the effectiveness of microbicides. An acidic pH is necessary for natural protection of the vagina against pathogens. Therefore, it is important to test the effectiveness of compounds in an acidic medium and in the presence of seminal plasma to probe their stability, safety and efficacy. 4.1.4. HIV-transmission inhibition assays Sexual transmission is one of the main routes of HIV transmission, mediated by exposure to infected cells or seminal fluid secretions in the mucosa. Currently, whether primary HIV infection is due to transmission of cell-free virus or cell-associated viruses is not well understood. The vaginal or rectal epitheliums are the first sites of deposition for cell-free or cell-associated viruses contained in the semen during sexual intercourse. The use of transwell permeable supports has been proposed as a model of these physiological conditions that can be used to study the potential of microbicide candidates to block the transmission of cell free or cell-associated HIV particles through the genital epithelium (Chonco et al., 2012; Gali et al., 2010a). This dual-chamber system consists of a tight epithelial cell layer or explants from vaginal, uterine cervical, endometrial or rectal mucosa grown on a transwell insert, and a basal chamber with PBMC. This model can be used to monitor whether treatment of the epithelial cells with the microbicide candidate prevents transmission of cell-free or cell-associated virus through the epithelial cell monolayer, simulating natural viral transmission in the vaginal/rectal epithelium. The infection of indicator cells, such as PBMC, in the lower chamber of the transwell is used to evaluate HIV transepithelial transmission. Moreover, transepithelial electrical resistance (TEER, Xxcm2) across the monolayer of epithelial cells can be monitored using a voltmeter and an electrode to investigate whether alterations of tight junction dynamics in the cell cultures are induced by treatment with the compounds. Several studies have shown the importance of using human vaginal or rectal explants to test the antiviral characteristics of the compounds in a physiologically relevant manner (Gali et al., 2010b; McElrath et al., 2010; Rohan et al., 2010). Interestingly, some nanoparticles have been already tested using these models. Polyanionic carbosilane dendrimers G3-S16 and G2-NF16 were able to prevent the transmission of cell-associated R5-HIV particles through transepithelial monolayers in transwell assays and they could block HIV internalization inside epithelial cells showing ability to protect the integrity of the monolayers (Vacas Cordoba et al., 2013). 4.1.5. Investigation of nanoparticle inhibitory mechanisms Following determination of the potency and biosafety of microbicide candidates, the next relevant step is to investigate the mechanisms by which these compounds inhibit infection. Ideal agents are those that block HIV-1 infection at early stages of the HIV-1 life cycle prior to viral integration, primarily by impeding HIV attachment, entry, fusion or reverse transcription. To date, the majority of nanotech-based microbicides act as HIV entry inhibitors. Due to their mechanism, they prevent the integration of the viral gen-

ome into the DNA of target cells, avoiding the primary infection and the establishment of a viral latent reservoir. Nevertheless, other nanocompounds with a late mode of action such as protease inhibitors (PI) could be also very efficacious blocking HIV sexual infection. PI are successfully used in HAART and present some advantages for their development as microbicide candidates. They exhibit a low adsorption in epithelium that may be advantageous in ensuring high local drug concentrations with minimal systemic exposure. Moreover, PI are the agents with the highest genetic barrier to resistance among ARV and due to their delayed antiviral activity, inhibiting maturation but not viral integration, an increased window of protection could be offered by these microbicide candidates (Herrera and Shattock, 2012), even exhibiting efficacy after post-coital application. Molecular modeling is an easy and potent tool to evaluate the targets on which the compounds exert their effects (Xue et al., 2014). For instance, molecular modeling was useful to establish the interaction between carbosilane dendrimer 2G-S16 and Gp120 and CD4 proteins (Chonco et al., 2012). Atomistic molecular dynamics simulations of complexes 2G-S16/CD4 and 2G-S16/Gp120 showed that dendrimer bound to cationic areas in those proteins such as Gp120-V3 loop region, ultimately blocking Gp120–CD4 interaction and thus the attachment step. Furthermore, molecular modeling can also be used to tailor and improve candidates in the search for more effective compounds. Other mechanistic studies, including cell-based and biochemical and enzymatic assays, can define the mode of action of the tested nanocompound. These include time-of-addition assays, in which a variety of antiretrovirals targeting different stages of the viral life cycle are compared to the activity of the tested compound at different time points, or gp120 capture ELISAs, which study the inhibitory activity of compounds on gp120–CD4 binding (Lara et al., 2010a).

604

4.1.6. Combination assays The combination of at least two families of ARV is more effective than the use of monotherapy for HIV treatment. Therefore, the combination of different compounds should be taken into consideration when designing new microbicide strategies (Anton, 2012; Balzarini and Schols, 2012). However, combination therapy has various significant disadvantages such as multiple drug–drug interactions, the potential development of multidrug resistances or the additive toxicity from the combination therapy leading to poor adherence. Nanotechnology can help to overcome these limitations. Nanomaterials have emerged as promising tools to increase the delivery of antiretroviral drugs or improve their pharmacokinetic and pharmacodynamic properties, certainly generating positive interactions. Polyanionic carbosilane dendrimers combined with different ARV, such as maraviroc or TFV, show synergistic activity against several laboratory and primary HIV isolates (Cordoba et al., 2013a; Sepulveda-Crespo et al., 2014; VacasCordoba et al., 2014). Efavirenz and saquinavir were loaded into Q3 PLGA nanoparticles in another combinatorial microbicide candidate study. The drug released from the nanoparticles was effective at inhibiting R5-tropic HIV in the human TZM-bl cell line (Chaowanachan et al., 2013). These combinatorial strategies have the potential to block HIV-1 infection at different stages of the HIV-1 life cycle, and may explain the positive interaction outcomes that are commonly observed. Moreover, targeting multiple stages within the HIV-1 replication cycle may decrease the risk of transmission of resistant viral isolates as well as reduce the development of resistance. Decreasing the amounts of antiviral compounds used in vivo might result in the reduction of adverse effects, limiting local toxicity and inflammation caused by topical application of the microbicide before sexual intercourse. Several software packages are available to analyze and statistically evalu-

636

Please cite this article in press as: Sánchez-Rodríguez, J., et al. Nanotech-derived topical microbicides for HIV prevention: The road to clinical development. Antiviral Res. (2014), http://dx.doi.org/10.1016/j.antiviral.2014.10.014

605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635

637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667

AVR 3537

No. of Pages 16, Model 5G

11 November 2014 J. Sánchez-Rodríguez et al. / Antiviral Research xxx (2014) xxx–xxx 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730

ate the outcome of combination assays (Buckheit and Lunsford, 2009; Chou and Talalay, 1984). 4.1.7. Broad spectrum activity of topical microbicides A desirable consideration in the development of new microbicides is not only effectiveness against diverse primary isolates, including clinical, drug resistant and multidrug resistant HIV viral strains, but also against other sexually transmitted pathogens, such as herpes simplex virus, hepatitis viruses, human papillomavirus and causative agents of bacterial and fungal vaginosis including Candida albicans, Neisseria gonorrhoeae or Gardnerella vaginalis. Therefore, it is crucial to evaluate the range of action of topical microbicide agents. It is especially important to develop effective microbicides against herpes simplex type 2 (HSV-2), the virus that causes genital herpes, HSV-2 infection is the most prevalent STD worldwide, with more than 500 million people infected and approximately 20 million new infections occurring each year. Moreover, genital herpes is an important co-factor for HIV-1 transmission; individuals infected with HSV-2 have a threefold higher risk of acquiring HIV-1 infection (Thurman and Doncel, 2012). Several mechanisms may account for this HSV-2-mediated facilitation of HIV-1 infection. HSV-2 infection causes ulceration of the genital mucosa, increasing the risk of HIV-1 acquisition. Furthermore, latent HSV2 infection increases the levels of CD4 and CCR5 expression on lymphocytes and macrophages and the levels of DC-SIGN on dendritic cells in the genital mucosa 5 6, increasing the probability of HIV infection (Rebbapragada et al., 2007; Sartori et al., 2011). As opposed to HIV, which primarily infects cells of the immune system, HSV-2 infects epithelial cells, neurons and other cell types. Therefore, distinct attachment and entry pathways are involved in HSV-2 entry into target cells. This attachment includes interactions between several glycoproteins in the viral envelope (gB, gD, gH, and gL) and a variety of cell surface molecules, including heparan sulfate moieties (Spear and Longnecker, 2003). Compounds capable of inhibiting this initial association and binding, as well as subsequent membrane fusion, could be potential microbicide candidates. The prediction algorithm and follow-up steps used to evaluate the antiviral activity of nanostructures against HSV-2 are the same as those used for HIV infection. However, several distinct in vitro and in vivo models have been successfully developed to evaluate inhibitory potency, such as the use of epithelial Vero cells for infection or plaque reduction assays for viral titration (Blaho et al., 2005). Some nanotech-derived architectures, such as membrane interacting peptide-functionalized dendrimers or zinc acetate gel, have been tested as potential anti-HSV-2 microbicides (Kenney et al., 2013; Luganini et al., 2011). Several nanoparticles have shown a broad spectrum of activity. For instance, phosphatidylcholine liposomes containing the microbicide octylglycerol were effective in vitro against HSV-1, HSV-2 and HIV-1 (Wang et al., 2012). SPL7013 dendrimer-based VivaGel demonstrated antiviral activity against HIV and HSV in animal models and antibacterial activity preventing bacterial vaginosis (Price et al., 2011), and silver nanoparticle-coated condoms were also able to inhibit HIV and HSV infection and demonstrated antibacterial and antifungal activities against several microorganism such as Escherichia coli, Staphylococcus aureus, Micrococcus luteus, Klebsiella pneumoniae, Candida tropicalis, Candida krusei, Candida glabrata, and Candida albicans (Mohammed Fayaz et al., 2012). 4.2. Ex vivo and in vivo evaluation: a proof-of-concept of compound efficacy prior to human clinical trials Because HIV replication is restricted to humans, and the level of biological complexity of vaginal and rectal transmission of HIV

11

cannot be mimicked in cell or tissue culture models, the research community has been searching for alternative in vivo study strategies for a long time. Sensitive animal models, that express the specific receptor and coreceptors that HIV-1 uses for attachment and entry in target cells (CD4, CCR5, or CXCR4), such as non-human primates and humanized mice (Brehm et al., 2010; Denton and Garcia, 2011; Veazey et al., 2012) play an important role in preclinical testing of microbicides, specially on safety, pharmacokinetic and efficacy testing. However, these animal models are not yet validated for US FDA (U.S. Food and Drug Administration) for predicting efficacy in humans (Administration, 2012). To date, the in vivo test based on the rabbit remains the only model recommended by the US FDA for the safety evaluation of vaginal products. The most widely used non-human primates models are macaques, including the rhesus (Macaca mulatta), pigtail (Macaca nemestrina), and cynomolgus (Macaca fascicularis) sub-species. These species have anatomical, immunological, and reproductive physiological properties broadly comparable to those of humans and they can be efficiently infected by different HIV strains. In case of rodents, their cells are non-permissive for the replication of HIV-1. Although rats, mice and rabbits, are certainly useful for preclinical safety studies of microbicide candidates, their reproductive cycles and tissues are quite different from those of humans or primates. Nevertheless, several humanized mouse models have been engineered recently to study HIV infection in vivo: (i) mice with single or multiple human transgenes; (ii) immunodeficient mice transplanted with human cells or tissues (healthy or diseased); (iii) immunodeficient mice transplanted with human CD34+ cells; or (iv) a combination of these three types. Currently, the most common model is the recently developed NOD/ SCIDcc / , Rag2 / cc / and blood, liver and thymus human (h-BLT) mouse model, which is reconstituted with human CD34 cells (NSG/BLT). In this model, immunodeficient mice are transplanted with human bone marrow (containing T cell progenitors), liver (for fetal immune development), and thymus (for T cell maturation) cells. Following engraftment, the mice are allowed to recover for 12 weeks prior to assessment for engraftment efficiency. These mice can completely reconstitute and self-renew all major human hematopoietic lineages (T and B cells, monocyte/ macrophage, dendritic and NK cells), essentially reproducing a functional human immune system that responds to X4- and R5tropic HIV-1 infection. The female reproductive tract of humanized BLT mice is reconstituted with functional human immune cells, rendering them susceptible to vaginal HIV-1 infection and allowing for testing of microbicide candidates. Recently, several laboratories have used this mouse model to evaluate the effectiveness of novel HIV microbicide candidates (Chateau et al., 2013; Denton et al., 2008, 2010, 2011; Di Fabio et al., 2003; Neff et al., 2011; Veselinovic et al., 2012; Wheeler et al., 2011). To test the ability of microbicides to prevent HIV infection, these studies focused on the evaluation of the presence of HIV particles in treated mice through different molecular techniques. Viral infection and viral loads are commonly assessed by Q-RT-PCR and western blot and flow cytometry is used to measure the levels of CD4 T cells in the peripheral blood of monitored mice. Although humanized mice are certainly useful, they present some obstacles, like the high cost and complexity of the surgical techniques required to generate them. Non-human primate models, due to their similarities to humans, remain the best option for safety and efficacy testing of new microbicide candidates. These in vivo models, together with ex vivo explant cultures, are also highly useful in evaluating the biosafety profiles of new nanobased microbicides in the final steps of their preclinical development, prior to toxicological analysis in humans. To study the impact of the compounds in the mucosal tissues, compounds must

Please cite this article in press as: Sánchez-Rodríguez, J., et al. Nanotech-derived topical microbicides for HIV prevention: The road to clinical development. Antiviral Res. (2014), http://dx.doi.org/10.1016/j.antiviral.2014.10.014

731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796

AVR 3537

No. of Pages 16, Model 5G

11 November 2014 12 797 798 799 800 801 802 803 804 805 806 807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825

J. Sánchez-Rodríguez et al. / Antiviral Research xxx (2014) xxx–xxx

be tested vaginally and/or rectally in order to evaluate potential toxic effects though direct observation of lesions in the vaginal or rectal area and the quantification of pro-inflammatory cytokines. Histopathological analysis of vaginal tissues can also be performed using semi-quantitative clinical scores to determine a vaginal irritation index (Eckstein et al., 1969). Epithelial lesions are evaluated based on the existence of hyperplasia or hyperkeratosis and ulceration, and/or inflammatory infiltrate in the epithelium. The inflammation process is characterized by an increased infiltration of lymphocytes compared to healthy mucosal tissues. Therefore, analysis of the lymphocyte score could help determine microbicide efficacy. Pro-inflammatory chemokines and cytokines plays an important role in HIV infection, owing to the recruitment and activation of CD4+ T cells to the mucosa (Fichorova et al., 2004). As mentioned previously, the concentration of inflammatory chemokines and cytokines in culture supernatants, serum samples or vaginal secretions is important in assessing the biosafety of microbicides. Furthermore, immunohistochemical studies on vaginal explants could be performed to study the levels of several biomarkers of cervicovaginal inflammation, such as IL-1, IL-6, TNFa, IL-8, macrophage inflammatory protein (MIP) or RANTES. Some nanoplatforms have been tested as topical microbicides in animal models. RNAi-mediated CCR5 silencing by LFA-1-targeted nanoparticles prevented HIV infection in BLT mice (Kim et al., 2010b). SPL7013 gel was protective against vaginal challenge with X4- and R5-tropic simian humanized immunodeficiency virus (SHIV) in a macaque model (Jiang et al., 2005). Moreover, activity against HSV-2 infection was demonstrated in mice and guinea pigs (Bernstein et al., 2003).

826

5. Future perspectives, obstacles and conclusions

827

Nanomedicine has provided new tools and rationales in the development and search for new and more powerful PrEP strategies, particularly in the design and engineering of more potent topical microbicides. The rational design and manipulation of diverse nanosystems, including dendrimers, liposomes and polymeric nanoparticles, has provided new microbicide candidates against HIV and other STDs with improved solubility, stability and pharmacokinetic and adherence properties. Furthermore, the combination of these nanoplatforms with other inhibitory drugs targeting different stages in the virus life cycle is essential for the development of more powerful and effective microbicides and for the improvement of existing ones. Since one of the most important problems in microbicide development is the lack of adherence in patients, a critical point to consider is the optimization of the nano-based microbicide formulation (Wong et al., 2014). Topical gel, intravaginal rings, films, tablets, ovules and other novel formulations such as female condoms coated with these nanostructures could be used. In each case, the formulation should be designed to maximize the activity, pharmacokinetics, safety, acceptability and adherence to the regimen of the different candidates. However, other limitations such as the adherence measurement or the comprehension of microbicide instructions for use in participating patients should be also improved to overcome adherence problems as reviewed in depth by Woodsong et al. (2013). In addition, novel routes of administration must be explored in order to discover novel nanotech-based prophylactic strategies against HIV. Although there are benefits and advantages of using nanotechnology in the development of novel microbicides to stop the spread of HIV worldwide, there are features and challenges that must be overcome in the future for successful translation of nano-based microbicides from in vitro and in vivo assays to clinical settings. These include the biocompatibility, safety, stability and cost of nanocompounds, as well as their scale-up for large-scale

828 829 830 831 832 833 834 835 836 837 838 839 840 841 842 843 844 845 846 847 848 849 850 851 852 853 854 855 856 857 858 859

production. More efficient and ecologically sustainable production methods or searching for active natural nanostructures like vegetable polymers, that can be large-scale produced in biofactories, could be essential points to reduce the production cost of nano-based microbicides. Plant systems offer unique advantages in comparison with conventional techniques (bacteria or yeast) or chemical synthesis such as lower production costs or an easy scale-up and purification process. These economic aspects are especially important because the main target of these products are the most vulnerable populations that live mainly in poor and developing countries, particularly in Africa. Because the majority of microbicides have failed in the last phases of human clinical trials, it is critical to begin the evaluation of topical microbicide candidates in conditions as similar as possible to physiological conditions that occur during sexual intercourse at the moment of HIV transmission. Pivotal points are the study of the impact of microbicides on vaginal and rectal fluid, on mucosal immune-cross-talk and the consequences of physical trauma during sex on the use of the microbicides. Currently, several nanosystems are under investigation as potential microbicides, with several on the threshold of clinical trials in humans. The encouraging results obtained in the CAPRISA 004 clinical trial and with VivaGel are proof-of-concepts that topical microbicides, and specifically nanotechnology-derived systems, can be effective approaches in the context of blocking HIV1 sexual transmission. These results support further clinical trials and in vivo research studies to develop a powerful strategy in the fight against the spread of the HIV/AIDS epidemic worldwide.

860

Acknowledgments

888

The authors do not have a commercial or other association that might pose a conflict of interest. This work was supported by grants provided through Fondo de Investigación Sanitaria (FIS) [grant number PI13/02016], Red Q4 Española de Investigación en SIDA (RIS) [grant numbers RD120017-0037, PT13/0010/0028 and RD12/0017/0029], CYTED 214RT0482, ‘‘Fundación para la Investigación y la Prevención del Sida en España’’ (FIPSE), Comunidad de Madrid [grant numbers, S-2010/BMD-2351 and S-2010/BMD-2332], ME&C (Ref. CTQ201123245). CIBER-BBN is an initiative funded by the VI National Q5 R&D&i Plan 2008-2011, Iniciativa Ingenio 2010, Consolider Program, CIBER Actions and financed by the Instituto de Salud Carlos III with assistance from the European Regional Development Fund.

889

References

902

Abdool Karim, S.S., Baxter, C., 2014. Microbicides for prevention of HIV infection: clinical efficacy trials. Curr. Top. Microbiol. Immunol. 383, 97–115. Abdool Karim, Q., Abdool Karim, S.S., Frohlich, J.A., Grobler, A.C., Baxter, C., Mansoor, L.E., Kharsany, A.B., Sibeko, S., Mlisana, K.P., Omar, Z., Gengiah, T.N., Maarschalk, S., Arulappan, N., Mlotshwa, M., Morris, L., Taylor, D., 2010. Effectiveness and safety of tenofovir gel, an antiretroviral microbicide, for the prevention of HIV infection in women. Science 329, 1168–1174. Administration, U.S.D.o.H.a.H.S.F.a.D., 2012. Guidance for industry: ‘‘vaginal microbicides: development for the prevention of HIV infection’’. Clin. Antimicrob. Amiji, M.M., Vyas, T.K., Shah, L.K., 2006. Role of nanotechnology in HIV/AIDS treatment: potential to overcome the viral reservoir challenge. Discov. Med. 6, 157–162. Anton, P.A., 2012. Future prospects and perspectives on microbicides. Curr. HIV Res. 10, 113–115. Asaftei, S., Huskens, D., Schols, D., 2012. HIV-1 X4 activities of polycationic ‘‘viologen’’ based dendrimers by interaction with the chemokine receptor CXCR4: study of structure–activity relationship. J. Med. Chem. 55, 10405– 10413. Baert, L., van ‘t Klooster, G., Dries, W., Francois, M., Wouters, A., Basstanie, E., Iterbeke, K., Stappers, F., Stevens, P., Schueller, L., Van Remoortere, P., Kraus, G., Wigerinck, P., Rosier, J., 2009. Development of a long-acting injectable formulation with nanoparticles of rilpivirine (TMC278) for HIV treatment. Eur. J. Pharm. Biopharm. 72, 502–508.

903 904 905 906 907 908 909 910 911 912 913 914 915 916 917 918 919 920 921 922 923 924 925 926

Please cite this article in press as: Sánchez-Rodríguez, J., et al. Nanotech-derived topical microbicides for HIV prevention: The road to clinical development. Antiviral Res. (2014), http://dx.doi.org/10.1016/j.antiviral.2014.10.014

861 862 863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887

890 891 892 893 894 895 896 897 898 899 900 901

AVR 3537

No. of Pages 16, Model 5G

11 November 2014 J. Sánchez-Rodríguez et al. / Antiviral Research xxx (2014) xxx–xxx 927 928 929 930 931 932 933 934 935 936 937 938 939 940 941 942 943 944 945 946 947 948 949 950 951 952 953 954 955 956 957 958 959 960 961 962 963 964 965 966 967 968 969 970 971 972 973 974 975 976 977 978 979 980 981 982 983 984 985 986 987 988 989 990 991 992 993 994 995 996 997 998 999 1000 1001 1002 1003 1004 1005 1006 1007 1008 1009 1010 1011

Ballagh, S.A., Baker, J.M., Henry, D.M., Archer, D.F., 2002. Safety of single daily use for one week of C31G HEC gel in women. Contraception 66, 369–375. Balzarini, J., Schols, D., 2012. Combination of antiretroviral drugs as microbicides. Curr. HIV Res. 10, 53–60. Balzarini, J., Van Damme, L., 2005. Intravaginal and intrarectal microbicides to prevent HIV infection. CMAJ 172, 461–464. Baranova, E.O., Shastina, N.S., Shvets, V.I., 2011. Polyanionic inhibitors of HIV adsorption. Bioorg. Khim. 37, 592–608. Bax, R., Douville, K., McCormick, D., Rosenberg, M., Higgins, J., Bowden, M., 2002. Microbicides – evaluating multiple formulations of C31G. Contraception 66, 365–368. Belletti, D., Tosi, G., Forni, F., Gamberini, M.C., Baraldi, C., Vandelli, M.A., Ruozi, B., 2012. Chemico-physical investigation of tenofovir loaded polymeric nanoparticles. Int. J. Pharm. 436, 753–763. Bernstein, D.I., Stanberry, L.R., Sacks, S., Ayisi, N.K., Gong, Y.H., Ireland, J., Mumper, R.J., Holan, G., Matthews, B., McCarthy, T., Bourne, N., 2003. Evaluations of unformulated and formulated dendrimer-based microbicide candidates in mouse and guinea pig models of genital herpes. Antimicrob. Agents Chemother. 47, 3784–3788. Bestman-Smith, J., Piret, J., Desormeaux, A., Tremblay, M.J., Omar, R.F., Bergeron, M.G., 2001. Sodium lauryl sulfate abrogates human immunodeficiency virus infectivity by affecting viral attachment. Antimicrob. Agents Chemother. 45, 2229–2237. Blaho, J.A., Morton, E.R., Yedowitz, J.C., 2005. Herpes simplex virus: propagation quantification and storage. Curr. Protoc. Microbiol. (Chapter 14, Unit 14E 11). Bourinbaiar, A.S., Fruhstorfer, E.C., 1996. The efficacy of nonoxynol-9 from an in vitro point of view. AIDS 10, 558–559. Brehm, M.A., Shultz, L.D., Greiner, D.L., 2010. Humanized mouse models to study human diseases. Curr. Opin. Endocrinol. Diabetes Obes. 17, 120–125. Briz, V., Serramia, M.J., Madrid, R., Hameau, A., Caminade, A.M., Majoral, J.P., MunozFernandez, M.A., 2012. Validation of a generation 4 phosphorus-containing polycationic dendrimer for gene delivery against HIV-1. Curr. Med. Chem. 19, 5044–5051. Buckheit Jr., R.W., Buckheit, K.W., 2012. An algorithm for the preclinical development of anti-HIV topical microbicides. Curr. HIV Res. 10, 97–104. Buckheit, R., Lunsford, R.D., 2009. In vitro performance and analysis of combination anti-infective evaluations. In: Mayers, D. (Ed.), Antimicrobial drug resistance. Humana Press, pp. 1135–1149. Buckheit Jr., R.W., Watson, K.M., Morrow, K.M., Ham, A.S., 2010. Development of topical microbicides to prevent the sexual transmission of HIV. Antiviral Res. 85, 142–158. Burello, E., Worth, A.P., 2011. QSAR modeling of nanomaterials. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 3, 298–306. Carlson, C., Hussain, S.M., Schrand, A.M., Braydich-Stolle, L.K., Hess, K.L., Jones, R.L., Schlager, J.J., 2008. Unique cellular interaction of silver nanoparticles: sizedependent generation of reactive oxygen species. J. Phys. Chem. B 112, 13608– 13619. Carter, A.J., Bourgeois, S., O’Brien, N., Abelsohn, K., Tharao, W., Greene, S., Margolese, S., Kaida, A., Sanchez, M., Palmer, A.K., Cescon, A., de Pokomandy, A., Loutfy, M.R., 2013. Women-specific HIV/AIDS services: identifying and defining the components of holistic service delivery for women living with HIV/AIDS. J. Int. AIDS Soc. 16, 17433. Chaowanachan, T., Krogstad, E., Ball, C., Woodrow, K.A., 2013. Drug synergy of tenofovir and nanoparticle-based antiretrovirals for HIV prophylaxis. PLoS ONE 8, e61416. Chateau, M.L., Denton, P.W., Swanson, M.D., McGowan, I., Garcia, J.V., 2013. Rectal transmission of transmitted/founder HIV-1 is efficiently prevented by topical 1% tenofovir in BLT humanized mice. PLoS ONE 8, e60024. Chen, B.A., Panther, L., Hoesley, C., et al., 2014. 21st Conference on Retroviruses and Opportunistic Infections (CROI). Abstract 41, (CROI) Safety and Pharmacokinetics/Pharmacodynamics of Dapivirine and Maraviroc Vaginal Rings, Boston, March 3–6. Chirenje, Z.M., Marrazzo, J., Parikh, U.M., 2010. Antiretroviral-based HIV prevention strategies for women. Expert Rev. Anti. Infect. Ther. 8, 1177–1186. Chonco, L., Pion, M., Vacas, E., Rasines, B., Maly, M., Serramia, M.J., Lopez-Fernandez, L., De la Mata, J., Alvarez, S., Gomez, R., Munoz-Fernandez, M.A., 2012. Carbosilane dendrimer nanotechnology outlines of the broad HIV blocker profile. J. Control. Release 161, 949–958. Chou, T.C., Talalay, P., 1984. Quantitative analysis of dose-effect relationships: the combined effects of multiple drugs or enzyme inhibitors. Adv. Enzyme Regul. 22, 27–55. Cordoba, E.V., Arnaiz, E., De La Mata, F.J., Gomez, R., Leal, M., Pion, M., MunozFernandez, M.A., 2013a. Synergistic activity of carbosilane dendrimers in combination with maraviroc against HIV in vitro. AIDS 27, 2053–2058. Cordoba, E.V., Bastida, H., Pion, M., Hameau, A., Ionov, M., Bryszewska, M., Caminade, A.M., Majoral, J.P., Munoz-Fernandez, M.A., 2013b. HIV-antigens charged on phosphorus dendrimers as tools for tolerogenic dendritic cellsbased immunotherapy. Curr. Med. Chem. (Epub ahead of print). Cordoba, E.V., Pion, M., Rasines, B., Filippini, D., Komber, H., Ionov, M., Bryszewska, M., Appelhans, D., Munoz-Fernandez, M.A., 2013c. Glycodendrimers as new tools in the search for effective anti-HIV DC-based immunotherapies. Nanomedicine 9, 972–984. Cu, Y., Saltzman, W.M., 2009. Controlled surface modification with poly(ethylene)glycol enhances diffusion of PLGA nanoparticles in human cervical mucus. Mol. Pharm. 6, 173–181.

13

das Neves, J., Amiji, M.M., Bahia, M.F., Sarmento, B., 2010. Nanotechnology-based systems for the treatment and prevention of HIV/AIDS. Adv. Drug Deliv. Rev. 62, 458–477. das Neves, J., Michiels, J., Arien, K.K., Vanham, G., Amiji, M.M., Bahia, M.F., Sarmento, B., 2012a. Polymeric nanoparticles affect the intracellular delivery, antiretroviral activity and cytotoxicity of the microbicide drug candidate dapivirine. Pharm. Res. 29, 1468–1484. das Neves, J., Rocha, C.M., Goncalves, M.P., Carrier, R.L., Amiji, M., Bahia, M.F., Sarmento, B., 2012b. Interactions of microbicide nanoparticles with a simulated vaginal fluid. Mol. Pharm. 9, 3347–3356. das Neves, J., Amiji, M., Bahia, M.F., Sarmento, B., 2013. Assessing the physical– chemical properties and stability of dapivirine-loaded polymeric nanoparticles. Int. J. Pharm. 456, 307–314. Date, A.A., Destache, C.J., 2013. A review of nanotechnological approaches for the prophylaxis of HIV/AIDS. Biomaterials 34, 6202–6228. De Clercq, E., 2012. Where rilpivirine meets with tenofovir, the start of a new antiHIV drug combination era. Biochem. Pharmacol. 84, 241–248. de Las Cuevas, N., Garcia-Gallego, S., Rasines, B., de la Mata, F.J., Guijarro, L.G., Munoz-Fernandez, M.A., Gomez, R., 2012. In vitro studies of water-stable cationic carbosilane dendrimers as delivery vehicles for gene therapy against HIV and hepatocarcinoma. Curr. Med. Chem. 19, 5052–5061. Denton, P.W., Garcia, J.V., 2011. Humanized mouse models of HIV infection. AIDS Rev. 13, 135–148. Denton, P.W., Estes, J.D., Sun, Z., Othieno, F.A., Wei, B.L., Wege, A.K., Powell, D.A., Payne, D., Haase, A.T., Garcia, J.V., 2008. Antiretroviral pre-exposure prophylaxis prevents vaginal transmission of HIV-1 in humanized BLT mice. PLoS Med. 5, e16. Denton, P.W., Krisko, J.F., Powell, D.A., Mathias, M., Kwak, Y.T., Martinez-Torres, F., Zou, W., Payne, D.A., Estes, J.D., Garcia, J.V., 2010. Systemic administration of antiretrovirals prior to exposure prevents rectal and intravenous HIV-1 transmission in humanized BLT mice. PLoS ONE 5, e8829. Denton, P.W., Othieno, F., Martinez-Torres, F., Zou, W., Krisko, J.F., Fleming, E., Zein, S., Powell, D.A., Wahl, A., Kwak, Y.T., Welch, B.D., Kay, M.S., Payne, D.A., Gallay, P., Appella, E., Estes, J.D., Lu, M., Garcia, J.V., 2011. One percent tenofovir applied topically to humanized BLT mice and used according to the CAPRISA 004 experimental design demonstrates partial protection from vaginal HIV infection, validating the BLT model for evaluation of new microbicide candidates. J. Virol. 85, 7582–7593. Dezzutti, C.S., Guenthner, P.C., Cummins Jr., J.E., Cabrera, T., Marshall, J.H., Dillberger, A., Lal, R.B., 2001. Cervical and prostate primary epithelial cells are not productively infected but sequester human immunodeficiency virus type 1. J. Infect. Dis. 183, 1204–1213. Di Fabio, S., Van Roey, J., Giannini, G., van den Mooter, G., Spada, M., Binelli, A., Pirillo, M.F., Germinario, E., Belardelli, F., de Bethune, M.P., Vella, S., 2003. Inhibition of vaginal transmission of HIV-1 in hu-SCID mice by the nonnucleoside reverse transcriptase inhibitor TMC120 in a gel formulation. AIDS 17, 1597–1604. Di Gianvincenzo, P., Marradi, M., Martinez-Avila, O.M., Bedoya, L.M., Alcami, J., Penades, S., 2010. Gold nanoparticles capped with sulfate-ended ligands as antiHIV agents. Bioorg. Med. Chem. Lett. 20, 2718–2721. Duncan, R., Izzo, L., 2005. Dendrimer biocompatibility and toxicity. Adv. Drug Deliv. Rev. 57, 2215–2237. Eckstein, P., Jackson, M.C., Millman, N., Sobrero, A.J., 1969. Comparison of vaginal tolerance tests of spermicidal preparations in rabbits and monkeys. J. Reprod. Fertil. 20, 85–93. Edwards, J.N., Morris, H.B., 1985. Langerhans’ cells and lymphocyte subsets in the female genital tract. Br. J. Obstet. Gynaecol. 92, 974–982. Emau, P., Tian, B., O’Keefe, B.R., Mori, T., McMahon, J.B., Palmer, K.E., Jiang, Y., Bekele, G., Tsai, C.C., 2007. Griffithsin, a potent HIV entry inhibitor, is an excellent candidate for anti-HIV microbicide. J. Med. Primatol. 36, 244–253. Ensign, L.M., Cone, R., Hanes, J., 2014. Nanoparticle-based drug delivery to the vagina: a review. J. Control. Release 190, 500–514. Fackler, O.T., Murooka, T.T., Imle, A., Mempel, T.R., 2014. Adding new dimensions: towards an integrative understanding of HIV-1 spread. Nat. Rev. Microbiol. 12, 563–574. Fahmy, T.M., Demento, S.L., Caplan, M.J., Mellman, I., Saltzman, W.M., 2008. Design opportunities for actively targeted nanoparticle vaccines. Nanomedicine (London) 3, 343–355. Farokhzad, O.C., 2008. Nanotechnology for drug delivery: the perfect partnership. Expert Opin. Drug Deliv. 5, 927–929. Feldblum, P.J., Adeiga, A., Bakare, R., Wevill, S., Lendvay, A., Obadaki, F., Olayemi, M.O., Wang, L., Nanda, K., Rountree, W., 2008. SAVVY vaginal gel (C31G) for prevention of HIV infection: a randomized controlled trial in Nigeria. PLoS ONE 3, e1474. Fetherston, S.M., Boyd, P., McCoy, C.F., McBride, M.C., Edwards, K.L., Ampofo, S., Malcolm, R.K., 2012. A silicone elastomer vaginal ring for HIV prevention containing two microbicides with different mechanisms of action. Eur. J. Pharm. Sci. 48, 406–415. Fichorova, R.N., Bajpai, M., Chandra, N., Hsiu, J.G., Spangler, M., Ratnam, V., Doncel, G.F., 2004. Interleukin (IL)-1, IL-6, and IL-8 predict mucosal toxicity of vaginal microbicidal contraceptives. Biol. Reprod. 71, 761–769. Friend, D.R., Kiser, P.F., 2013. Assessment of topical microbicides to prevent HIV-1 transmission: concepts, testing, lessons learned. Antiviral Res. 99, 391–400. Galan, M., Sanchez Rodriguez, J., Jimenez, J.L., Relloso, M., Maly, M., de la Mata, F.J., Munoz-Fernandez, M.A., Gomez, R., 2014. Synthesis of new anionic carbosilane

Please cite this article in press as: Sánchez-Rodríguez, J., et al. Nanotech-derived topical microbicides for HIV prevention: The road to clinical development. Antiviral Res. (2014), http://dx.doi.org/10.1016/j.antiviral.2014.10.014

1012 1013 1014 1015 1016 1017 1018 1019 1020 1021 1022 1023 1024 1025 1026 1027 1028 1029 1030 1031 1032 1033 1034 1035 1036 1037 1038 1039 1040 1041 1042 1043 1044 1045 1046 1047 1048 1049 1050 1051 1052 1053 1054 1055 1056 1057 1058 1059 1060 1061 1062 1063 1064 1065 1066 1067 1068 1069 1070 1071 1072 1073 1074 1075 1076 1077 1078 1079 1080 1081 1082 1083 1084 1085 1086 1087 1088 1089 1090 1091 1092 1093 1094 1095 1096

AVR 3537

No. of Pages 16, Model 5G

11 November 2014 14 1097 1098 1099 1100 1101 1102 1103 1104 1105 1106 1107 1108 1109 1110 1111 1112 1113 1114 1115 1116 1117 1118 1119 1120 1121 1122 1123 1124 1125 1126 1127 1128 1129 1130 1131 1132 1133 1134 1135 1136 1137 1138 1139 1140 1141 1142 1143 1144 1145 1146 1147 1148 1149 1150 1151 1152 1153 1154 1155 1156 1157 1158 1159 1160 1161 1162 1163 1164 1165 1166 1167 1168 1169 1170 1171 1172 1173 1174 1175 1176 1177 1178 1179 1180 1181

J. Sánchez-Rodríguez et al. / Antiviral Research xxx (2014) xxx–xxx

dendrimers via thiol-ene chemistry and their antiviral behaviour. Org. Biomol. Chem. 12, 3222–3237. Gali, Y., Arien, K.K., Praet, M., Van den Bergh, R., Temmerman, M., Delezay, O., Vanham, G., 2010a. Development of an in vitro dual-chamber model of the female genital tract as a screening tool for epithelial toxicity. J. Virol. Methods 165, 186–197. Gali, Y., Delezay, O., Brouwers, J., Addad, N., Augustijns, P., Bourlet, T., HamzehCognasse, H., Arien, K.K., Pozzetto, B., Vanham, G., 2010b. In vitro evaluation of viability, integrity, and inflammation in genital epithelia upon exposure to pharmaceutical excipients and candidate microbicides. Antimicrob. Agents Chemother. 54, 5105–5114. Garcia-Gallego, S., Rodriguez, J.S., Jimenez, J.L., Cangiotti, M., Ottaviani, M.F., MunozFernandez, M.A., Gomez, R., de la Mata, F.J., 2012. Polyanionic N-donor ligands as chelating agents in transition metal complexes: synthesis, structural characterization and antiviral properties against HIV. Dalton Trans. 41, 6488–6499. Garcia-Vallejo, J.J., Koning, N., Ambrosini, M., Kalay, H., Vuist, I., Sarrami-Forooshani, R., Geijtenbeek, T.B., van Kooyk, Y., 2013. Glycodendrimers prevent HIV transmission via DC-SIGN on dendritic cells. Int. Immunol. 25, 221–233. Gengiah, T.N., Baxter, C., Mansoor, L.E., Kharsany, A.B., Abdool Karim, S.S., 2012. A drug evaluation of 1% tenofovir gel and tenofovir disoproxil fumarate tablets for the prevention of HIV infection. Expert Opin. Investig. Drugs 21, 695–715. Gibson, R.M., Arts, E.J., 2012. Past, present, and future of entry inhibitors as HIV microbicides. Curr. HIV Res. 10, 19–26. Gong, E., Matthews, B., McCarthy, T., Chu, J., Holan, G., Raff, J., Sacks, S., 2005. Evaluation of dendrimer SPL7013, a lead microbicide candidate against herpes simplex viruses. Antiviral Res. 68, 139–146. Granich, R., Crowley, S., Vitoria, M., Lo, Y.R., Souteyrand, Y., Dye, C., Gilks, C., Guerma, T., De Cock, K.M., Williams, B., 2010. Highly active antiretroviral treatment for the prevention of HIV transmission. J. Int. AIDS Soc. 13, 1. Grivel, J.C., Shattock, R.J., Margolis, L.B., 2011. Selective transmission of R5 HIV-1 variants: where is the gatekeeper? J. Transl. Med. 9 (Suppl. 1), S6. Gunn, J., Zhang, M., 2010. Polyblend nanofibers for biomedical applications: perspectives and challenges. Trends Biotechnol. 28, 189–197. Gupta, S.K., 2013. Clinical use of vaginal or rectally applied microbicides in patients suffering from HIV/AIDS. HIV AIDS 5, 295–307. Gupta, P., Lackman-Smith, C., Snyder, B., Ratner, D., Rohan, L.C., Patton, D., Ramratnam, B., Cole, A.M., 2013. Antiviral activity of retrocyclin RC-101, a candidate microbicide against cell-associated HIV-1. AIDS Res. Hum. Retroviruses 29, 391–396. Ham, A.S., Cost, M.R., Sassi, A.B., Dezzutti, C.S., Rohan, L.C., 2009. Targeted delivery of PSC-RANTES for HIV-1 prevention using biodegradable nanoparticles. Pharm. Res. 26, 502–511. Herrera, C., Shattock, R.J., 2012. Potential use of protease inhibitors as vaginal and colorectal microbicides. Curr. HIV Res. 10, 42–52. Hillier, S., 2013. Overview of Products in Development-Topical. Forum for Coolborative HIV Research. In: Meeting on Future of PrEP and Microbicides, Washington. Hladik, F., McElrath, M.J., 2008. Setting the stage: host invasion by HIV. Nat. Rev. Immunol. 8, 447–457. Hladik, F., Sakchalathorn, P., Ballweber, L., Lentz, G., Fialkow, M., Eschenbach, D., McElrath, M.J., 2007. Initial events in establishing vaginal entry and infection by human immunodeficiency virus type-1. Immunity 26, 257–270. Huang, C., Soenen, S.J., van Gulck, E., Vanham, G., Rejman, J., Van Calenbergh, S., Vervaet, C., Coenye, T., Verstraelen, H., Temmerman, M., Demeester, J., De Smedt, S.C., 2012. Electrospun cellulose acetate phthalate fibers for semen induced anti-HIV vaginal drug delivery. Biomaterials 33, 962–969. Huskens, D., Vermeire, K., Profy, A.T., Schols, D., 2009. The candidate sulfonated microbicide, PRO 2000, has potential multiple mechanisms of action against HIV-1. Antiviral Res. 84, 38–47. Jiang, Y.H., Emau, P., Cairns, J.S., Flanary, L., Morton, W.R., McCarthy, T.D., Tsai, C.C., 2005. SPL7013 gel as a topical microbicide for prevention of vaginal transmission of SHIV89.6P in macaques. AIDS Res. Hum. Retroviruses 21, 207–213. Jimenez, J.L., Clemente, M.I., Weber, N.D., Sanchez, J., Ortega, P., de la Mata, F.J., Gomez, R., Garcia, D., Lopez-Fernandez, L.A., Munoz-Fernandez, M.A., 2010. Carbosilane dendrimers to transfect human astrocytes with small interfering RNA targeting human immunodeficiency virus. BioDrugs 24, 331–343. Keller, M.J., Mesquita, P.M., Torres, N.M., Cho, S., Shust, G., Madan, R.P., Cohen, H.W., Petrie, J., Ford, T., Soto-Torres, L., Profy, A.T., Herold, B.C., 2010. Postcoital bioavailability and antiviral activity of 0.5% PRO 2000 gel: implications for future microbicide clinical trials.. PLoS ONE 5, e8781. Kenney, J., Aravantinou, M., Singer, R., Hsu, M., Rodriguez, A., Kizima, L., Abraham, C.J., Menon, R., Seidor, S., Chudolij, A., Gettie, A., Blanchard, J., Lifson, J.D., Piatak Jr., M., Fernandez-Romero, J.A., Zydowsky, T.M., Robbiani, M., 2011. An antiretroviral/zinc combination gel provides 24 hours of complete protection against vaginal SHIV infection in macaques. PLoS ONE 6, e15835. Kenney, J., Rodriguez, A., Kizima, L., Seidor, S., Menon, R., Jean-Pierre, N., Pugach, P., Levendosky, K., Derby, N., Gettie, A., Blanchard, J., Piatak Jr., M., Lifson, J.D., Paglini, G., Zydowsky, T.M., Robbiani, M., Fernandez Romero, J.A., 2013. A modified zinc acetate gel, a potential non-antiretroviral microbicide, is safe and effective against simian-human immunodeficiency virus and herpes simplex virus 2 infection in vivo. Antimicrob. Agents Chemother. 57, 4001–4009. Kim, P.S., Read, S.W., 2010. Nanotechnology and HIV: potential applications for treatment and prevention. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2, 693–702.

Kim, B.Y., Rutka, J.T., Chan, W.C., 2010a. Nanomedicine. N. Engl. J. Med. 363, 2434– 2443. Kim, S.S., Peer, D., Kumar, P., Subramanya, S., Wu, H., Asthana, D., Habiro, K., Yang, Y.G., Manjunath, N., Shimaoka, M., Shankar, P., 2010b. RNAi-mediated CCR5 silencing by LFA-1-targeted nanoparticles prevents HIV infection in BLT mice. Mol. Ther. 18, 370–376. Klasse, P.J., Shattock, R., Moore, J.P., 2008. Antiretroviral drug-based microbicides to prevent HIV-1 sexual transmission. Annu. Rev. Med. 59, 455–471. Krebs, F.C., Miller, S.R., Malamud, D., Howett, M.K., Wigdahl, B., 1999. Inactivation of human immunodeficiency virus type 1 by nonoxynol-9, C31G, or an alkyl sulfate, sodium dodecyl sulfate. Antiviral Res. 43, 157–173. Lara, H.H., Ayala-Nunez, N.V., Ixtepan-Turrent, L., Rodriguez-Padilla, C., 2010a. Mode of antiviral action of silver nanoparticles against HIV-1. J. Nanobiotechnol. 8, 1. Lara, H.H., Ixtepan-Turrent, L., Garza-Trevino, E.N., Rodriguez-Padilla, C., 2010b. PVP-coated silver nanoparticles block the transmission of cell-free and cellassociated HIV-1 in human cervical culture. J. Nanobiotechnol. 8, 15. Lazniewska, J., Milowska, K., Gabryelak, T., 2012. Dendrimers – revolutionary drugs for infectious diseases. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 4, 469–491. Lederman, M.M., Veazey, R.S., Offord, R., Mosier, D.E., Dufour, J., Mefford, M., Piatak Jr., M., Lifson, J.D., Salkowitz, J.R., Rodriguez, B., Blauvelt, A., Hartley, O., 2004. Prevention of vaginal SHIV transmission in rhesus macaques through inhibition of CCR5. Science 306, 485–487. Luganini, A., Nicoletto, S.F., Pizzuto, L., Pirri, G., Giuliani, A., Landolfo, S., Gribaudo, G., 2011. Inhibition of herpes simplex virus type 1 and type 2 infections by peptide-derivatized dendrimers. Antimicrob. Agents Chemother. 55, 3231– 3239. Maeda, K., Das, D., Nakata, H., Mitsuya, H., 2012. CCR5 inhibitors: emergence, success, and challenges. Expert Opin. Emerg. Drugs 17, 135–145. Mahajan, S.D., Law, W.C., Aalinkeel, R., Reynolds, J., Nair, B.B., Yong, K.T., Roy, I., Prasad, P.N., Schwartz, S.A., 2012. Nanoparticle-mediated targeted delivery of antiretrovirals to the brain. Methods Enzymol. 509, 41–60. Malavia, N.K., Zurakowski, D., Schroeder, A., Princiotto, A.M., Laury, A.R., Barash, H.E., Sodroski, J., Langer, R., Madani, N., Kohane, D.S., 2011. Liposomes for HIV prophylaxis. Biomaterials 32, 8663–8668. Mallipeddi, R., Rohan, L.C., 2010. Progress in antiretroviral drug delivery using nanotechnology. Int. J. Nanomed. 5, 533–547. Mamo, T., Moseman, E.A., Kolishetti, N., Salvador-Morales, C., Shi, J., Kuritzkes, D.R., Langer, R., von Andrian, U., Farokhzad, O.C., 2010. Emerging nanotechnology approaches for HIV/AIDS treatment and prevention. Nanomedicine (London) 5, 269–285. Martin, D.H., 2012. The microbiota of the vagina and its influence on women’s health and disease. Am. J. Med. Sci. 343, 2–9. Mauck, C.K., Weiner, D.H., Creinin, M.D., Barnhart, K.T., Callahan, M.M., Bax, R., 2004. A randomized phase I vaginal safety study of three concentrations of C31G vs. Extra strength gynol II. Contraception 70, 233–240. Mbopi-Keou, F.X., Trottier, S., Omar, R.F., Nkele, N.N., Fokoua, S., Mbu, E.R., Domingo, M.C., Giguere, J.F., Piret, J., Mwatha, A., Masse, B., Bergeron, M.G., 2010. A randomized, double-blind, placebo-controlled phase II extended safety study of two invisible condom formulations in Cameroonian women. Contraception 81, 79–85. McElrath, M.J., Ballweber, L., Terker, A., Kreger, A., Sakchalathorn, P., Robinson, B., Fialkow, M., Lentz, G., Hladik, F., 2010. Ex vivo comparison of microbicide efficacies for preventing HIV-1 genomic integration in intraepithelial vaginal cells. Antimicrob. Agents Chemother. 54, 763–772. McGowan, I., Gomez, K., Bruder, K., Febo, I., Chen, B.A., Richardson, B.A., Husnik, M., Livant, E., Price, C., Jacobson, C., 2011. Phase 1 randomized trial of the vaginal safety and acceptability of SPL7013 gel (VivaGel) in sexually active young women (MTN-004). AIDS 25, 1057–1064. Meng, J., Sturgis, T.F., Youan, B.B., 2011. Engineering tenofovir loaded chitosan nanoparticles to maximize microbicide mucoadhesion. Eur. J. Pharm. Sci. 44, 57–67. Mertenskoetter, T., Kaptur, P.E., 2011. Update on microbicide research and development – seeking new HIV prevention tools for women. Eur. J. Med. Res. 16, 1–6. Microbicides, I.P.F., 2012. The Ring Study. Mohammed Fayaz, A., Ao, Z., Girilal, M., Chen, L., Xiao, X., Kalaichelvan, P., Yao, X., 2012. Inactivation of microbial infectiousness by silver nanoparticles-coated condom: a new approach to inhibit HIV- and HSV-transmitted infection. Int. J. Nanomed. 7, 5007–5018. Morrow, K.M., Hendrix, C., 2010. Clinical evaluation of microbicide formulations. Antiviral Res. 88 (Suppl. 1), 40–46. Munier, C.M., Andersen, C.R., Kelleher, A.D., 2011. HIV vaccines: progress to date. Drugs 71, 387–414. Navath, R.S., Menjoge, A.R., Dai, H., Romero, R., Kannan, S., Kannan, R.M., 2011. Injectable PAMAM dendrimer-PEG hydrogels for the treatment of genital infections: formulation and in vitro and in vivo evaluation. Mol. Pharm. 8, 1209– 1223. Neff, C.P., Kurisu, T., Ndolo, T., Fox, K., Akkina, R., 2011. A topical microbicide gel formulation of CCR5 antagonist maraviroc prevents HIV-1 vaginal transmission in humanized RAG-hu mice. PLoS ONE 6, e20209. Network, M.T., 2010. Understanding the Results of CAPRISA 004. Network, M.T., 2012. Phase III Trial of Dapivirine Ring Begins in Africa: ASPIRE Testing New HIV Prevention Approach for Women.

Please cite this article in press as: Sánchez-Rodríguez, J., et al. Nanotech-derived topical microbicides for HIV prevention: The road to clinical development. Antiviral Res. (2014), http://dx.doi.org/10.1016/j.antiviral.2014.10.014

1182 1183 1184 1185 1186 1187 1188 1189 1190 1191 1192 1193 1194 1195 1196 1197 1198 1199 1200 1201 1202 1203 1204 1205 1206 1207 1208 1209 1210 1211 1212 1213 1214 1215 1216 1217 1218 1219 1220 1221 1222 1223 1224 1225 1226 1227 1228 1229 1230 1231 1232 1233 1234 1235 1236 1237 1238 1239 1240 1241 1242 1243 1244 1245 1246 1247 1248 1249 1250 1251 1252 1253 1254 1255 1256 1257 1258 1259 1260 1261 1262 1263 1264 1265 1266

AVR 3537

No. of Pages 16, Model 5G

11 November 2014 J. Sánchez-Rodríguez et al. / Antiviral Research xxx (2014) xxx–xxx 1267 1268 1269 1270 1271 1272 1273 1274 1275 1276 1277 1278 1279 1280 1281 1282 1283 1284 1285 1286 1287 1288 1289 1290 1291 1292 1293 1294 1295 1296 1297 1298 1299 1300 1301 1302 1303 1304 1305 1306 1307 1308 1309 1310 1311 1312 1313 1314 1315 1316 1317 1318 1319 1320 1321 1322 1323 1324 1325 1326 1327 1328 1329 1330 1331 1332 1333 1334 1335 1336 1337 1338 1339 1340 1341 1342 1343 1344 1345 1346 1347 1348 1349 1350 1351 1352

Network, M.T., 2013a. Phase 1 Evaluation of Coitus on the Pharmacokinetics and Pharmacodynamics of Tenofovir 1% Gel Following Pericoital or Daily Gel Dosing. Network, M.T., 2013b. A Phase 2 Randomized Sequence Open Label Expanded Safety and Acceptability Study of Oral Emtricitabine/Tenofovir Disoproxil Fumarate Tablet and Rectally-Applied Tenofovir Reduced-Glycerin 1% Gel. O’Loughlin, J., Millwood, I.Y., McDonald, H.M., Price, C.F., Kaldor, J.M., Paull, J.R., 2010. Safety, tolerability, and pharmacokinetics of SPL7013 gel (VivaGel): a dose ranging, phase I study. Sex. Transm. Dis. 37, 100–104. Omar, R.F., Bergeron, M.G., 2011. The future of microbicides. Int. J. Infect. Dis. 15, 656–660. Padian, N.S., Buve, A., Balkus, J., Serwadda, D., Cates Jr., W., 2008. Biomedical interventions to prevent HIV infection: evidence, challenges, and way forward. Lancet 372, 585–599. Parboosing, R., Maguire, G.E., Govender, P., Kruger, H.G., 2012. Nanotechnology and the Treatment of HIV Infection. Viruses 4, 488–520. Perise-Barrios, A.J., Jimenez, J.L., Dominguez-Soto, A., de la Mata, F.J., Corbi, A.L., Gomez, R., Munoz-Fernandez, M.A., 2014. Carbosilane dendrimers as gene delivery agents for the treatment of HIV infection. J. Control. Release 184, 51– 57. Piret, J., Desormeaux, A., Bergeron, M.G., 2002. Sodium lauryl sulfate, a microbicide effective against enveloped and nonenveloped viruses. Curr. Drug Targets 3, 17– 30. Pirrone, V., Wigdahl, B., Krebs, F.C., 2011. The rise and fall of polyanionic inhibitors of the human immunodeficiency virus type 1. Antiviral Res. 90, 168–182. Price, C.F., Tyssen, D., Sonza, S., Davie, A., Evans, S., Lewis, G.R., Xia, S., Spelman, T., Hodsman, P., Moench, T.R., Humberstone, A., Paull, J.R., Tachedjian, G., 2011. SPL7013 Gel (VivaGel(R)) retains potent HIV-1 and HSV-2 inhibitory activity following vaginal administration in humans. PLoS ONE 6, e24095. Quinn, T.C., 2007. Circumcision and HIV transmission. Curr. Opin. Infect. Dis. 20, 33– 38. Rasines, B., Sanchez-Nieves, J., Maiolo, M., Maly, M., Chonco, L., Jimenez, J.L., MunozFernandez, M.A., de la Mata, F.J., Gomez, R., 2012. Synthesis, structure and molecular modelling of anionic carbosilane dendrimers. Dalton Trans. 41, 12733–12748. Rebbapragada, A., Wachihi, C., Pettengell, C., Sunderji, S., Huibner, S., Jaoko, W., Ball, B., Fowke, K., Mazzulli, T., Plummer, F.A., Kaul, R., 2007. Negative mucosal synergy between Herpes simplex type 2 and HIV in the female genital tract. AIDS 21, 589–598. Reina, J.J., Bernardi, A., Clerici, M., Rojo, J., 2010. HIV microbicides: state-of-the-art and new perspectives on the development of entry inhibitors. Future Med. Chem. 2, 1141–1159. Roddy, R.E., Zekeng, L., Ryan, K.A., Tamoufe, U., Weir, S.S., Wong, E.L., 1998. A controlled trial of nonoxynol 9 film to reduce male-to-female transmission of sexually transmitted diseases. N. Engl. J. Med. 339, 504–510. Rohan, L.C., Moncla, B.J., Kunjara Na Ayudhya, R.P., Cost, M., Huang, Y., Gai, F., Billitto, N., Lynam, J.D., Pryke, K., Graebing, P., Hopkins, N., Rooney, J.F., Friend, D., Dezzutti, C.S., 2010. In vitro and ex vivo testing of tenofovir shows it is effective as an HIV-1 microbicide. PLoS ONE 5, e9310. Rohan, L.C., Yang, H., Wang, L., 2013. Rectal pre-exposure prophylaxis (PrEP). Antiviral Res. 100 (Suppl.), 17–24. Rosa Borges, A., Wieczorek, L., Johnson, B., Benesi, A.J., Brown, B.K., Kensinger, R.D., Krebs, F.C., Wigdahl, B., Blumenthal, R., Puri, A., McCutchan, F.E., Birx, D.L., Polonis, V.R., Schengrund, C.L., 2010. Multivalent dendrimeric compounds containing carbohydrates expressed on immune cells inhibit infection by primary isolates of HIV-1. Virology 408, 80–88. Sanchez-Nieves, J., Fransen, P., Pulido, D., Lorente, R., Munoz-Fernandez, M.A., Albericio, F., Royo, M., Gomez, R., de la Mata, F.J., 2014. Amphiphilic cationic carbosilane-PEG dendrimers: synthesis and applications in gene therapy. Eur. J. Med. Chem. 76, 43–52. Sartori, E., Calistri, A., Salata, C., Del Vecchio, C., Palu, G., Parolin, C., 2011. Herpes simplex virus type 2 infection increases human immunodeficiency virus type 1 entry into human primary macrophages. Virol. J. 8, 166. Sepulveda-Crespo, D., Lorente, R., Leal, M., Gomez, R., De La Mata, F.J., Jimenez, J.L., Munoz-Fernandez, M.A., 2014. Synergistic activity profile of carbosilane dendrimer G2-STE16 in combination with other dendrimers and antiretrovirals as topical anti-HIV-1 microbicide. Nanomedicine 10, 609–618. Shattock, R.J., Rosenberg, Z., 2012. Microbicides: topical prevention against HIV. Cold Spring Harb. Perspect. Med. 2, a007385. Siccardi, M., Martin, P., McDonald, T.O., Liptrott, N.J., Giardiello, M., Rannard, S., Owen, A., 2013. Nanomedicines for HIV therapy. Ther. Deliv. 4, 153–156. Spear, P.G., Longnecker, R., 2003. Herpesvirus entry: an update. J. Virol. 77, 10179– 10185. Stefanidou, M., Herrera, C., Armanasco, N., Shattock, R.J., 2012. Saquinavir inhibits early events associated with establishment of HIV-1 infection: potential role for protease inhibitors in prevention. Antimicrob. Agents Chemother. 56, 4381– 4390. Steinbach, J.M., Weller, C.E., Booth, C.J., Saltzman, W.M., 2012. Polymer nanoparticles encapsulating siRNA for treatment of HSV-2 genital infection. J. Control. Release 162, 102–110. Stephenson, J., 2011. Studies probe new anti-HIV strategy, long-term success of prevention methods. JAMA 305, 1397–1399. Svenson, S., Tomalia, D.A., 2005. Dendrimers in biomedical applications – reflections on the field. Adv. Drug Deliv. Rev. 57, 2106–2129. Tang, B.C., Dawson, M., Lai, S.K., Wang, Y.Y., Suk, J.S., Yang, M., Zeitlin, P., Boyle, M.P., Fu, J., Hanes, J., 2009. Biodegradable polymer nanoparticles that rapidly

15

penetrate the human mucus barrier. Proc. Natl. Acad. Sci. U.S.A. 106, 19268– 19273. Tao, W., Richards, C., Hamer, D., 2008. Enhancement of HIV infection by cellulose sulfate. AIDS Res. Hum. Retroviruses 24, 925–929. Telwatte, S., Moore, K., Johnson, A., Tyssen, D., Sterjovski, J., Aldunate, M., Gorry, P.R., Ramsland, P.A., Lewis, G.R., Paull, J.R., Sonza, S., Tachedjian, G., 2011. Virucidal activity of the dendrimer microbicide SPL7013 against HIV-1. Antiviral Res. 90, 195–199. Thiagarajan, G., Greish, K., Ghandehari, H., 2013. Charge affects the oral toxicity of poly(amidoamine) dendrimers. Eur. J. Pharm. Biopharm. 84, 330–334. Thompson, K.A., Malamud, D., Storey, B.T., 1996. Assessment of the anti-microbial agent C31G as a spermicide: comparison with nonoxynol-9. Contraception 53, 313–318. Thurman, A.R., Doncel, G.F., 2012. Herpes simplex virus and HIV: genital infection synergy and novel approaches to dual prevention. Int. J. STD AIDS 23, 613– 619. Tsai, C.C., Emau, P., Jiang, Y., Tian, B., Morton, W.R., Gustafson, K.R., Boyd, M.R., 2003. Cyanovirin-N gel as a topical microbicide prevents rectal transmission of SHIV89.6P in macaques. AIDS Res. Hum. Retroviruses 19, 535–541. Turpin, J.A., 2011. Topical microbicides to prevent the transmission of HIV: formulation gaps and challenges. Drug Deliv. Transl. Res. 1, 194–200. Tyssen, D., Henderson, S.A., Johnson, A., Sterjovski, J., Moore, K., La, J., Zanin, M., Sonza, S., Karellas, P., Giannis, M.P., Krippner, G., Wesselingh, S., McCarthy, T., Gorry, P.R., Ramsland, P.A., Cone, R., Paull, J.R., Lewis, G.R., Tachedjian, G., 2010. Structure activity relationship of dendrimer microbicides with dual action antiviral activity. PLoS ONE 5, e12309. UNAIDS, J.U.N.P.o.H.A., 2013. Global Report: UNAIDS Report on the Global AIDS Epidemic 2013. Vacas Cordoba, E., Arnaiz, E., Relloso, M., Sanchez-Torres, C., Garcia, F., PerezAlvarez, L., Gomez, R., de la Mata, F.J., Pion, M., Munoz-Fernandez, M.A., 2013. Development of sulphated and naphthylsulphonated carbosilane dendrimers as topical microbicides to prevent HIV-1 sexual transmission. AIDS 27, 1219– 1229. Vacas-Cordoba, E., Galan, M., de la Mata, F.J., Gomez, R., Pion, M., Munoz-Fernandez, M.A., 2014. Enhanced activity of carbosilane dendrimers against HIV when combined with reverse transcriptase inhibitor drugs: searching for more potent microbicides. Int. J. Nanomed. 9, 3591–3600. Van Damme, L., Ramjee, G., Alary, M., Vuylsteke, B., Chandeying, V., Rees, H., Sirivongrangson, P., Mukenge-Tshibaka, L., Ettiegne-Traore, V., Uaheowitchai, C., Karim, S.S., Masse, B., Perriens, J., Laga, M., 2002. Effectiveness of COL-1492, a nonoxynol-9 vaginal gel, on HIV-1 transmission in female sex workers: a randomised controlled trial. Lancet 360, 971–977. van Marle, G., Gill, M.J., Kolodka, D., McManus, L., Grant, T., Church, D.L., 2007. Compartmentalization of the gut viral reservoir in HIV-1 infected patients. Retrovirology 4, 87. Veazey, R.S., Ketas, T.J., Dufour, J., Moroney-Rasmussen, T., Green, L.C., Klasse, P.J., Moore, J.P., 2010. Protection of rhesus macaques from vaginal infection by vaginally delivered maraviroc, an inhibitor of HIV-1 entry via the CCR5 coreceptor. J. Infect. Dis. 202, 739–744. Veazey, R.S., Shattock, R.J., Klasse, P.J., 2012. Animal models for microbicide studies. Curr. HIV Res. 10, 79–87. Veselinovic, M., Neff, C.P., Mulder, L.R., Akkina, R., 2012. Topical gel formulation of broadly neutralizing anti-HIV-1 monoclonal antibody VRC01 confers protection against HIV-1 vaginal challenge in a humanized mouse model. Virology 432, 505–510. Vijayakumar, S., Ganesan, S., 2012. Gold nanoparticles as an HIV entry inhibitor. Curr. HIV Res. 10, 643–646. Wang, L., Sassi, A.B., Patton, D., Isaacs, C., Moncla, B.J., Gupta, P., Rohan, L.C., 2012. Development of a liposome microbicide formulation for vaginal delivery of octylglycerol for HIV prevention. Drug Dev. Ind. Pharm. 38, 995–1007. Weber, N., Ortega, P., Clemente, M.I., Shcharbin, D., Bryszewska, M., de la Mata, F.J., Gomez, R., Munoz-Fernandez, M.A., 2008. Characterization of carbosilane dendrimers as effective carriers of siRNA to HIV-infected lymphocytes. J. Control. Release 132, 55–64. Weber, N.D., Merkel, O.M., Kissel, T., Munoz-Fernandez, M.A., 2012. PEGylated poly(ethylene imine) copolymer-delivered siRNA inhibits HIV replication in vitro. J. Control. Release 157, 55–63. Wheeler, L.A., Trifonova, R., Vrbanac, V., Basar, E., McKernan, S., Xu, Z., Seung, E., Deruaz, M., Dudek, T., Einarsson, J.I., Yang, L., Allen, T.M., Luster, A.D., Tager, A.M., Dykxhoorn, D.M., Lieberman, J., 2011. Inhibition of HIV transmission in human cervicovaginal explants and humanized mice using CD4 aptamer-siRNA chimeras. J. Clin. Invest. 121, 2401–2412. Witvrouw, M., Fikkert, V., Pluymers, W., Matthews, B., Mardel, K., Schols, D., Raff, J., Debyser, Z., De Clercq, E., Holan, G., Pannecouque, C., 2000. Polyanionic (i.e., polysulfonate) dendrimers can inhibit the replication of human immunodeficiency virus by interfering with both virus adsorption and later steps (reverse transcriptase/integrase) in the virus replicative cycle. Mol. Pharmacol. 58, 1100–1108. Wong, T.W., Dhanawat, M., Rathbone, M.J., 2014. Vaginal drug delivery: strategies and concerns in polymeric nanoparticle development. Expert Opin. Drug Deliv. 11, 1419–1434. Woodsong, C., MacQueen, K., Amico, K.R., Friedland, B., Gafos, M., Mansoor, L., Tolley, E., McCormack, S., 2013. Microbicide clinical trial adherence: insights for introduction. J. Int. AIDS Soc. 16, 18505.

Please cite this article in press as: Sánchez-Rodríguez, J., et al. Nanotech-derived topical microbicides for HIV prevention: The road to clinical development. Antiviral Res. (2014), http://dx.doi.org/10.1016/j.antiviral.2014.10.014

1353 1354 1355 1356 1357 1358 1359 1360 1361 1362 1363 1364 1365 1366 1367 1368 1369 1370 1371 1372 1373 1374 1375 1376 1377 1378 1379 1380 1381 1382 1383 1384 1385 1386 1387 1388 1389 1390 1391 1392 1393 1394 1395 1396 1397 1398 1399 1400 1401 1402 1403 1404 1405 1406 1407 1408 1409 1410 1411 1412 1413 1414 1415 1416 1417 1418 1419 1420 1421 1422 1423 1424 1425 1426 1427 1428 1429 1430 1431 1432 1433 1434 1435 1436 1437 1438

AVR 3537

No. of Pages 16, Model 5G

11 November 2014 16 1439 1440 1441 1442 1443 1444

J. Sánchez-Rodríguez et al. / Antiviral Research xxx (2014) xxx–xxx

Xiong, S., George, S., Yu, H., Damoiseaux, R., France, B., Ng, K.W., Loo, J.S., 2013. Size influences the cytotoxicity of poly (lactic-co-glycolic acid) (PLGA) and titanium dioxide (TiO(2)) nanoparticles. Arch. Toxicol. 87, 1075–1086. Xue, W., Liu, H., Yao, X., 2014. Molecular modeling study on the allosteric inhibition mechanism of HIV-1 integrase by LEDGF/p75 binding site inhibitors. PLoS ONE 9, e90799.

Zhang, L., He, T., Talal, A., Wang, G., Frankel, S.S., Ho, D.D., 1998. In vivo distribution of the human immunodeficiency virus/simian immunodeficiency virus coreceptors: CXCR4, CCR3, and CCR5. J. Virol. 72, 5035–5045. Zhang, T., Sturgis, T.F., Youan, B.B., 2011. PH-responsive nanoparticles releasing tenofovir intended for the prevention of HIV transmission. Eur. J. Pharm. Biopharm. 79, 526–536.

1445 1446 1447 1448 1449 1450 1451

Please cite this article in press as: Sánchez-Rodríguez, J., et al. Nanotech-derived topical microbicides for HIV prevention: The road to clinical development. Antiviral Res. (2014), http://dx.doi.org/10.1016/j.antiviral.2014.10.014

Nanotech-derived topical microbicides for HIV prevention: the road to clinical development.

More than three decades since its discovery, HIV infection remains one of the most aggressive epidemics worldwide, with more than 35 million people in...
2MB Sizes 0 Downloads 6 Views