Human immunodeficiency virus (HIV) is an enveloped, single-stranded RNA retrovirus belonging to the genus Lentivirus within the family Retroviridae, with two main types—HIV-1, responsible for the global pandemic, and HIV-2, primarily confined to West Africa.[1][2] The virus targets CD4+ T lymphocytes, macrophages, and dendritic cells by binding to CD4 receptors and co-receptors such as CCR5 or CXCR4, leading to viral entry, reverse transcription of its RNA genome into DNA, integration into the host genome, and subsequent production of new virions that progressively deplete immune cells.[2][3] Untreated, HIV causes acquired immunodeficiency syndrome (AIDS), characterized by severe opportunistic infections and cancers due to profound immunosuppression, though antiretroviral therapy (ART) can suppress replication to undetectable levels, preventing progression to AIDS and transmission.[2][4]HIV originated from multiple zoonotic transmissions of simian immunodeficiency viruses (SIVs) from African primates to humans, with HIV-1 group M—the dominant pandemic strain—emerging from SIVcpz in chimpanzees (Pan troglodytes troglodytes) in Central Africa around the early 20th century, likely via bushmeat hunting.[5][6] HIV-2 arose independently from SIVsmm in sooty mangabeys in West Africa.[5] Genetic and phylogenetic evidence strongly supports natural cross-species jumps rather than laboratory origins, with viral diversification occurring in human populations before widespread recognition of AIDS cases in the 1980s.[6][5] The virus spreads primarily through unprotected sexual contact, sharing contaminated needles, blood transfusions with unscreened blood, and mother-to-child transmission during pregnancy, birth, or breastfeeding, with transmission efficiency varying by viral load and route.[7][3]Isolated in 1983 by Luc Montagnier and Françoise Barré-Sinoussi at the Pasteur Institute from a patient with lymphadenopathy, HIV was soon propagated and characterized by Robert Gallo's team at the National Cancer Institute, confirming its role as the causative agent of AIDS through fulfillment of adapted Koch's postulates in human disease models.[8][9] A dispute over credit and virus sourcing led to U.S.-France agreements and the 2008 Nobel Prize in Physiology or Medicine awarded solely to the French team, excluding Gallo despite his contributions to HIV cultivation, testing, and etiology proof.[9][8] By 2024, approximately 40.8 million people lived with HIV globally, with sub-Saharan Africa bearing over two-thirds of cases; new infections have declined but plateaued short of elimination targets, while ART access has averted millions of AIDS deaths, transforming HIV into a manageable chronic condition for adherent patients.[10][4] Challenges persist in drug resistance, adherence, and disparities in prevention like pre-exposure prophylaxis efficacy against high-risk behaviors.[11][4]
Biology
Classification and Structure
Human immunodeficiency virus (HIV) belongs to the family Retroviridae, subfamily Orthoretrovirinae, and genus Lentivirus within the order Ortervirales.[12][13] Two types exist: HIV-1, which accounts for over 95% of global infections and originated from simian immunodeficiency virus (SIV) in chimpanzees (SIVcpz), and HIV-2, derived from SIV in sooty mangabeys (SIVsmm) and primarily circulating in West Africa.[14][15] HIV-1 comprises groups M (predominant worldwide), N, O, and P, with group M featuring subtypes A–D, F–H, J, K, and circulating recombinant forms (CRFs).[15] HIV-2 includes subtypes A–H, with A and B most prevalent; HIV-2 exhibits lower transmissibility and slower progression to AIDS than HIV-1.[14]The mature HIV virion measures 100–120 nm in diameter and assumes a roughly spherical shape with a lipidenvelope acquired from the host cell membrane.[17][18] Embedded in the envelope are ~72 trimers of the envelope glycoprotein complex, each comprising an external gp120 subunit for receptor binding and a transmembrane gp41 subunit for membranefusion.[18][19] The envelope underlies a matrix layer of p17 protein (MA), which supports virion assembly and links to the conical capsid formed by ~1,500–2,000 p24 molecules (CA).[20][21]The capsid encases two copies of positive-sense, single-stranded RNAgenome (~9.7 kb for HIV-1), bound by nucleocapsid proteins p7 (NC) that maintain dimerization and packaging signals.[20] Essential enzymes—reverse transcriptase (p66/p51 heterodimer), integrase (p32), and protease (p10)—reside within or near the capsid, enabling reverse transcription, genomeintegration, and polyprotein cleavage, respectively.[20] Accessory proteins like Vif, Vpr, Vpu (HIV-1) or Vpx (HIV-2), and Nef incorporate into the virion to counter host restrictions and enhance infectivity.[18] HIV-1 and HIV-2 share this core architecture, though HIV-2 encodes Vpx instead of Vpu, influencing early replication differences.[14]
Genome and Replication Cycle
The genome of human immunodeficiency virus (HIV) consists of two identical strands of single-stranded, positive-sense RNA, each approximately 9.7 kilobases (kb) in length, forming a diploid structure characteristic of retroviruses.[22] This RNA genome features a 5' cap and 3' poly-A tail, and is flanked by long terminal repeats (LTRs) of about 600 nucleotides each, which contain regulatory elements essential for viral transcription and integration.[22] The HIV-1 genome, the predominant form responsible for the global pandemic, encodes nine genes that produce 15 proteins through alternative splicing and ribosomal frameshifting: the structural genes gag (encoding matrix, capsid, and nucleocapsid proteins), pol (protease, reverse transcriptase, and integrase), and env (surface glycoprotein gp120 and transmembrane gp41); and accessory/regulatory genes tat (transactivation), rev (export of unspliced RNA), nef (downregulation of CD4 and MHC-I), vif (counteracts host restriction factors), vpr (nuclear import and cell cycle arrest), and vpu (enhances virion release and degrades CD4).[23] In HIV-2, vpu is replaced by vpx, which degrades SAMHD1 to promote replication in myeloid cells, though both viruses share core replication mechanisms.[24]HIV replication begins with attachment to the host cell surface via the viral envelopeglycoprotein gp120 binding to the CD4 receptor on target cells, followed by interaction with co-receptors such as CCR5 or CXCR4, enabling membrane fusion and release of the viral capsid into the cytoplasm.[25] The RNA genome is then reverse-transcribed by the virion-associated reverse transcriptase enzyme into double-stranded DNA within the partially uncoated capsid, a process that occurs in the cytoplasm and generates long terminal repeats at both ends of the proviral DNA.[26] This pre-integration complex, including integrase, is transported to the nucleus where integrase catalyzes the insertion of viral DNA into the host chromosome, forming a provirus that persists latently or actively transcribes.[25]Transcription of the provirus by host RNA polymerase II produces full-length viral RNA transcripts, which serve dual roles as genomic RNA for new virions and messenger RNA for protein synthesis; regulatory proteins Tat and Rev enhance this process by promoting elongation and nuclear export of unspliced/incompletely spliced transcripts, respectively.[26] Translated Gag and Gag-Pol polyproteins assemble at the plasma membrane with Env glycoproteins and two copies of genomic RNA, facilitated by accessory proteins like Vif and Nef; the immature virion buds from the cell, acquiring a host-derived lipid envelope.[25] Post-budding, viral protease cleaves Gag and Gag-Pol precursors, enabling maturation into infectious particles capable of initiating a new cycle, with each cycle lasting 1-2 days and producing thousands of progeny virions per infected cell.[27] This reverse transcription-dependent lifecycle distinguishes HIV as a lentivirus, allowing integration into non-dividing cells like macrophages and resting T-lymphocytes, contributing to chronic persistence.[28]
Tropism and Pathogenesis
HIV-1 primarily targets cells expressing the CD4 receptor, including helper T lymphocytes (CD4+ T cells), macrophages, and dendritic cells, with entry requiring interaction with a chemokine co-receptor, either CCR5 or CXCR4.[29][30] Strains utilizing CCR5 (R5-tropic or macrophage-tropic) predominate during early infection, efficiently infecting memory CD4+ T cells and macrophages in mucosal tissues, while those using CXCR4 (X4-tropic or T-tropic) emerge later in disease progression and preferentially target naive CD4+ T cells.[31][32] This shift in tropism correlates with advancing immunodeficiency, as X4-tropic variants associate with faster CD4+ T cell decline and higher viral loads.[33]Pathogenesis involves progressive depletion of CD4+ T cells, driven by direct viral cytopathic effects in infected cells, including pyroptosis—an inflammatory form of programmed cell death triggered by abortive infection—and bystander killing via immune activation and apoptosis.[34][35] Early after infection, massive CD4+ T cell loss occurs in gut-associated lymphoid tissues, where CCR5-expressing cells are abundant, disrupting mucosal immunity and enabling microbial translocation that fuels chronic inflammation.[36][37] Infected macrophages and dendritic cells contribute by serving as reservoirs for persistent viral production and by impairing antigen presentation, further exacerbating T cell dysfunction without equivalent depletion in these myeloid lineages.[38][39]Overall, CD4+ T cell counts typically decline from a pre-infection median of around 1,000 cells/μL to below 200 cells/μL over 5–10 years in untreated individuals, marking the transition to AIDS with susceptibility to opportunistic infections due to impaired cellular immunity.[40] Multiple mechanisms interplay, including Fas-independent direct killing and LFA-1/ICAM-1-mediated enhancements of viral spread and cell death, underscoring that pathogenesis extends beyond mere viral replication to systemic immune dysregulation.[41][42]
Transmission and Epidemiology
Modes of Transmission
HIV is transmitted exclusively through direct contact with specific infectious body fluids—blood, semen (including pre-seminal fluid), rectal fluids, vaginal fluids, and breast milk—from a person with detectable HIV viremia entering the mucous membranes, bloodstream, or open wounds of an uninfected individual.[43][44] Transmission requires sufficient viral load in the fluid and a portal of entry, such as mucosal tears, abrasions, or vascular access; the virus does not survive long outside the body and is not viable in saliva, urine, tears, sweat, or feces under typical exposure conditions due to low concentrations and natural inhibitory substances.[43][45]Sexual contact represents the predominant mode of HIV transmission worldwide, occurring via unprotected vaginal or anal intercourse when infected fluids contact genital or rectal mucosa.[43] Per-act transmission risks, derived from meta-analyses of cohort studies, are highest for receptive anal intercourse at 138 infections per 10,000 exposures (1.38%), followed by insertive anal intercourse at 11 per 10,000 (0.11%), receptive vaginal intercourse at 8 per 10,000 (0.08%), and insertive vaginal intercourse at 4 per 10,000 (0.04%).[46][47] These probabilities increase with higher viral loads in the infected partner, concurrent sexually transmitted infections causing mucosal inflammation, and lack of circumcision in insertive male partners; oral sex confers negligible risk, with no confirmed transmissions in large serodiscordant couple studies absent blood exposure.[43][48]Parenteral transmission via shared injection equipment among people who inject drugs introduces infected blood directly into the circulation, with per-sharing risks estimated at 63 per 10,000 acts (0.63%) based on observational data adjusting for viral load and syringe contamination volume.[47] Accidental needlestick injuries in healthcare settings carry a lower risk of approximately 23 per 10,000 exposures (0.23%), mitigated by immediate post-exposure prophylaxis.[47]Mother-to-child transmission (MTCT) occurs vertically during pregnancy (intrauterine), labor and delivery (peripartum), or breastfeeding (postnatal), with cumulative risks without antiretroviral intervention ranging from 15% to 45% depending on maternal viral load, mode of delivery, and breastfeeding duration—approximately 5-10% intrauterine, 10-20% peripartum, and 12-15% per year of exclusive breastfeeding.[49][50] Transmission via unscreened blood transfusions or contaminated blood products was historically near-certain (over 90% seroconversion rate) but has been reduced to below 1 in 1.5 million units transfused in screened systems through donor deferral, nucleic acid testing, and serologic assays implemented since the 1980s.[51][52]HIV is not transmitted through casual or household contact, as evidenced by longitudinal studies of over 300 exposed household members showing zero infections attributable to non-sexual, non-parenteral sharing of utensils, toilets, or saliva, with upper confidence limits excluding meaningful risk.[53][54] Rare case reports of other routes, such as female-to-female sexual contact or mucous membrane splashes, lack virologic confirmation and are confounded by unacknowledged high-risk behaviors.[55][45]
Risk Factors and Behavioral Determinants
Unprotected receptive anal intercourse represents the highest-risk sexual behavior for HIV acquisition, with a per-act transmission probability estimated at 1.38% (138 per 10,000 exposures) when the infected partner has detectable viral load.[48] Insertive anal intercourse carries a lower but still elevated risk of 0.11% (11 per 10,000 exposures).[48] Unprotected vaginal intercourse poses comparatively reduced risks, at 0.08% (8 per 10,000) for receptive penile-vaginal contact and 0.04% (4 per 10,000) for insertive, based on pooled data from cohort studies and meta-analyses of serodiscordant couples.[47]Oral sex confers negligible per-act risk, typically below 0.04% even without barriers.[48] These probabilities derive from empirical observations of transmission events relative to exposure acts, adjusted for factors like viral load but excluding pre-exposure prophylaxis use.[47]Sharing contaminated needles or syringes during injection drug use elevates risk substantially, with a per-act probability of approximately 0.63% (63 per 10,000 shares), reflecting direct bloodstream exposure to high viral concentrations.[48] This behavioral pathway accounts for a significant proportion of non-sexual transmissions, particularly in populations with frequent injection practices and limited access to sterile equipment. Multiple injections per session compound cumulative exposure, amplifying overall incidence independent of sexual behaviors.[56]Several modifiable factors intensify transmission efficiency beyond baseline per-act risks. Acute HIV infection in the source partner boosts infectivity by 7- to 26-fold due to peak viremia, while chronic high viral loads (>50,000 copies/mL) similarly elevate odds.[57] Coexisting sexually transmitted infections, such as genital ulcers or inflammation, increase mucosal susceptibility or viral shedding, raising risk by 2- to 5-fold across sexual acts.[58] Substance use, including alcohol intoxication or stimulants, correlates with impaired decision-making, leading to higher rates of unprotected sex or needle sharing; for instance, methamphetamine use among men who have sex with men (MSM) associates with 2- to 3-fold increased odds of unprotected anal intercourse.[58][59]Behavioral patterns drive disparities in HIV prevalence through cumulative exposures rather than isolated acts. Frequent partner concurrency or serial monogamy with unprotected sex heightens network-level spread, as each additional partner multiplies potential transmission opportunities.[60] Among MSM, repeated unprotected receptive anal intercourse explains disproportionate incidence, with U.S. data showing this group accounting for 69% of new diagnoses despite comprising 2-4% of the malepopulation, attributable to the behavior's inherent efficiency and prevalence within sexual networks.[59] Injection drug users exhibit elevated risks from both needle-sharing practices and overlapping sexual behaviors, such as trading sex for drugs, which further interconnect transmission routes.[56] These determinants underscore that HIV epidemiology reflects voluntary high-risk choices, modifiable through consistent barrier use or harm reduction, rather than unavoidable exposures.[61]
Global Prevalence and Demographic Patterns
As of 2024, approximately 40.8 million people [37.0–45.6 million] were living with HIV globally.[10] This includes 39.4 million [35.7–44.0 million] adults aged 15 years and older and 1.4 million [1.1–1.8 million] children aged 0–14 years.[10] Women and girls comprised 53% of all people living with HIV.[10] An estimated 1.3 million [1.0–1.7 million] individuals acquired HIV in 2024, reflecting a decline from prior peaks but stagnation short of elimination targets.[10]The epidemic exhibits stark regional disparities, with sub-Saharan Africa bearing the heaviest burden: eastern and southern Africa alone accounted for 20.8 million people living with HIV in 2023, comprising over half the global total, while western and central Africa added 5.1 million. Adult HIV prevalence in this region often exceeds 3%, with rates above 20% in high-burden countries such as Eswatini, Botswana, Lesotho, and Zimbabwe.[62] In contrast, Asia and the Pacific reported around 6.7 million cases, Latin America about 2.4 million, and Western and Central Europe and North America fewer than 2.5 million combined in 2023 estimates. Eastern Europe, Central Asia, the Middle East, and North Africa have seen rising infections in recent years, driven by injection drug use and other factors.[63]Demographic patterns reveal gender imbalances varying by transmission context. Globally, 45% of new 2024 infections occurred among women and girls, but this rose to 63% in sub-Saharan Africa, where heterosexual transmission predominates and biological and social factors—such as higher viral loads during acute infection in men and age-disparate partnerships—contribute to elevated acquisition rates among adolescent females.[10] Approximately 4,000 adolescent girls and young women aged 15–24 acquired HIV weekly in 2024, with 3,300 of these in sub-Saharan Africa.[10] In regions with lower overall prevalence, such as Western Europe, North America, and parts of Asia, men account for the majority of cases, largely due to disproportionate impact on men who have sex with men (MSM), who face HIV prevalence rates of about 7.6%.[10]Age distributions concentrate infections among young and middle-aged adults, with peak incidence in the 15–49 age group reflecting sexual debut, partnering behaviors, and cumulative exposure risks.[64] Children under 15 represent about 3.4% of global cases, primarily from mother-to-child transmission, though prevention of mother-to-child transmission programs have reduced new pediatric infections to 120,000 [82,000–170,000] in 2024.[64] Key populations, including MSM, people who inject drugs (prevalence ~7.1%), female sex workers (~2.7%), and transgender individuals (~8.5%), experience infection rates far exceeding general populations, accounting for a substantial share of new cases even in low-prevalence settings.[10]
HIV testing methods detect either host antibodies produced in response to infection, viral antigens such as p24, or the virus's genetic material via nucleic acid amplification.[65] The Centers for Disease Control and Prevention (CDC) recommends an initial screening with a fourth-generation antigen/antibodyimmunoassay, which identifies both HIV-1/2 antibodies and the p24 antigen, followed by confirmatory testing to distinguish true positives from false results.[66] This approach minimizes the diagnostic window period—the time between infection and detectable markers—to 18–45 days for laboratory-based antigen/antibody tests on venous blood.[67]Antibody-only tests, such as enzyme-linked immunosorbent assays (ELISA) or rapid immunochromatographic tests, identify IgM and IgG antibodies to HIV but have a longer window period of up to 90 days, as seroconversion may take 3 weeks to 3 months post-exposure.[68] These tests exhibit high sensitivity (>99.9%) and specificity (>99%), though false-positive rates can reach approximately 1.5% in low-prevalence settings due to cross-reactivity with other conditions, necessitating Western blot or nucleic acid confirmation.[69] Fourth-generation combination tests reduce the window by detecting p24 antigen earlier, shortening it by 4–5 days compared to third-generation antibody assays, with overall accuracy exceeding 99.9% for negative results.[70][71]Nucleic acid tests (NAT), including quantitative polymerase chain reaction (qPCR) for HIV RNA, enable the earliest detection at 10–33 days post-exposure and are essential for diagnosing acute infections or confirming indeterminate results in high-risk scenarios like occupational exposure.[65]NAT sensitivity approaches 99.9% for viral loads above 50 copies/mL, but false negatives occur at very low viral loads during early eclipse phases, and costs limit routine screening use to blood donation or infant diagnosis.[72] False-positive NAT results are rare (<0.1%) but have been reported in contexts like certain gene therapies due to sequence cross-reactivity.[73]
Test Type
Detects
Window Period (Days Post-Exposure)
Primary Use
Sensitivity/Specificity
Antibody-only (e.g., ELISA, rapid)
HIV-1/2 IgM/IgG
23–90
Screening in low-risk settings
>99.9% / >99%[74]
Antigen/Antibody (4th gen)
p24 Ag + antibodies
18–45 (lab); 18–90 (rapid)
Initial universal screening
>99.9% / >99.5%[75]
NAT (qPCR RNA)
Viral RNA
10–33
Acute diagnosis, confirmation, infants
~99.9% (at >50 copies/mL) / >99.9%[72]
Rapid point-of-care tests, approved since the early 1990s, provide results in 20–30 minutes using fingerstick blood or oral fluid, with self-testing kits available since 2012 for home use, though they retain a 90-day window for antibody detection.[76][77] All positive screens require follow-up, as unconfirmed positives yield more false results in populations with HIV prevalence below 1%, per Bayes' theorem considerations on pretest probability.[78] The first HIV antibody test emerged in 1985 for blood screening, evolving to fourth-generation assays by 2010, which now form the basis of CDC's algorithm updated in 2014 for acute HIV detection.[79][80]
Staging and Monitoring
HIV staging assesses disease progression based on clinical symptoms and immunological markers, primarily CD4 T-cell counts, to guide prognosis and management decisions. The Centers for Disease Control and Prevention (CDC) employs a dual classification system revised in 1993, combining clinical categories with CD4-based immunologic stages. Clinical category A denotes acute or asymptomaticinfection or persistent generalized lymphadenopathy; category B includes symptomatic conditions not meeting AIDS criteria, such as bacillary angiomatosis or oral hairy leukoplakia; and category C encompasses AIDS-defining illnesses like Pneumocystis pneumonia or esophageal candidiasis. Immunologic stages are stratified by CD4 count: stage 1 (>500 cells/mm³), stage 2 (200–499 cells/mm³), and stage 3 (<200 cells/mm³, defining AIDS regardless of symptoms).[81][82]The World Health Organization (WHO) uses a clinical staging system for resource-limited settings, classifying adults into four stages without requiring CD4 counts. Stage 1 is asymptomatic or limited to persistent generalized lymphadenopathy; stage 2 includes moderate unexplained symptoms like recurrent respiratory infections or herpes zoster; stage 3 involves severe conditions such as unexplained weight loss (>10% body weight), chronic diarrhea, or pulmonary tuberculosis; and stage 4 comprises AIDS-defining illnesses like extrapulmonary tuberculosis or Kaposi's sarcoma. WHO also incorporates immunological classification: stage 1 (CD4 >500 cells/mm³), stage 2 (200–499 cells/mm³), stage 3 (50–199 cells/mm³), and stage 4 (<50 cells/mm³). Advanced HIV disease is defined as CD4 <200 cells/mm³ or WHO stage 3/4.[83][84][85]Monitoring involves serial laboratory assessments to track viral replication, immune function, and treatment efficacy, enabling early detection of progression or antiretroviral therapy (ART) failure. Plasma HIV-1 RNA (viral load) quantifies viremia, with levels >100,000 copies/mL predicting faster progression in untreated individuals; suppression below 50 copies/mL on ART indicates success. For untreated patients, viral load and CD4 counts are checked every 3–6 months to assess disease advancement.[86][87]In ART recipients, guidelines recommend viral load testing at ART initiation, at 4–6 weeks, 3 months, then every 3–6 months if suppressed; intervals may extend to 6 months for stable patients with CD4 >200 cells/mm³. CD4 monitoring occurs every 3–6 months initially or if counts remain <300–500 cells/mm³, but becomes optional after 2 years of suppression with counts >500 cells/mm³, as it adds limited prognostic value beyond viral load. Genotypic resistance testing is advised for virologic failure (confirmed viral load >200 copies/mL), and adherence assessments complement labs to distinguish non-adherence from resistance. Routine monitoring also screens for comorbidities like hepatitis or tuberculosis, particularly in advanced stages.[86][88][89]
Clinical Course
Acute and Chronic Phases
The acute phase of HIV infection, occurring 2 to 4 weeks after initial viralexposure, manifests as acute retroviral syndrome in 40 to 90 percent of cases.[58][90] Symptoms include fever, maculopapular rash, sore throat, lymphadenopathy, myalgia, and fatigue, resembling influenza or mononucleosis, and typically resolve within 1 to 4 weeks.[91][92] During this period, plasma HIV RNA levels peak at 10^6 to 10^7 copies per milliliter around day 21 post-infection, reflecting massive viral replication and dissemination to lymphoid tissues.[93]CD4 T-cell counts drop transiently by 100 to 300 cells per microliter before partial recovery, while seroconversion to detectable anti-HIV antibodies occurs within 3 to 12 weeks.[92] This phase carries elevated transmission risk due to high viremia in blood and mucosal secretions.[92]The chronic phase, also termed clinical latency, follows resolution of acute symptoms and persists for a median of 10 years without antiretroviral therapy, though duration varies widely among individuals.[94][95] Viral replication continues at a lower set-point level, typically 10^3 to 10^4 copies per milliliter after the initial decline from peak, insufficient to cause overt symptoms in most but sufficient for ongoing CD4 T-cell depletion.[96] Patients often remain asymptomatic, with possible persistent generalized lymphadenopathy as the sole sign.[97]CD4 counts decline at an average rate of 50 to 100 cells per microliter per year, influenced by factors including baseline viral load and host genetics, leading to progression to AIDS when counts fall below 200 cells per microliter.[98][99] Approximately 25 percent of untreated individuals progress rapidly within 2 years, 50 percent over 2 to 10 years, and 25 percent more slowly beyond 10 years.[99] With modern antiretroviral therapy, chronicinfection can be managed indefinitely, preventing progression and restoring immune function.[94]
Opportunistic Infections and Complications
Opportunistic infections in HIV primarily arise due to progressive depletion of CD4+ T lymphocytes, impairing cell-mediated immunity and increasing susceptibility to pathogens that are typically controlled in immunocompetent hosts; these infections often define the onset of acquired immunodeficiencysyndrome (AIDS) when CD4 counts drop below 200 cells/μL.[100] Common manifestations include respiratory, central nervous system (CNS), and disseminated infections, with historical incidence rates exceeding 50% in untreated advanced cases prior to widespread antiretroviral therapy (ART) adoption.[101] Prophylaxis and early ART have reduced their frequency, but they remain prevalent in late presenters or resource-limited settings, where overall OI incidence can reach 7-13 per 100 person-years.[102][103]Pneumocystis jirovecii pneumonia (PCP), formerly Pneumocystis carinii, is among the most frequent OIs, characterized by subacute progressive dyspnea, fever, nonproductive cough, and hypoxemia, with radiographic findings of bilateral interstitial infiltrates; it occurs at CD4 counts typically under 200 cells/μL and was a leading cause of death in the pre-ART era, affecting up to 40% of AIDS patients without prophylaxis.[104]Tuberculosis (TB), particularly reactivation of latent Mycobacterium tuberculosis, presents with pulmonary or extrapulmonary disease at higher rates in HIV, with incidence densities up to 16% in co-endemic areas and often atypical radiographic patterns like lower lobe involvement or miliary spread.[100] Oropharyngeal and esophageal candidiasis, caused by Candida species, manifests as painful white plaques or dysphagia and signals CD4 decline below 200-300 cells/μL, recurring in over 90% of untreated patients.[101]CNS opportunistic infections include toxoplasmic encephalitis from Toxoplasma gondii, the most common brain mass lesion in AIDS (incidence 10-50% without prophylaxis), featuring ring-enhancing lesions on imaging, headache, focal deficits, and fever at CD4 <100 cells/μL.[105] Cryptococcal meningitis, due to Cryptococcus neoformans, causes subacute headache, altered mental status, and high intracranial pressure, with global incidence historically at 5-10% in advanced HIV before fluconazole prophylaxis became standard.[101] Cytomegalovirus (CMV) retinitis leads to vision-threatening retinal necrosis in 20-40% of untreated AIDS patients with CD4 <50 cells/μL, presenting as painless visual field loss.[105] Disseminated histoplasmosis and Mycobacterium avium complex (MAC) bacteremia occur at very low CD4 counts (<50 cells/μL), with MAC causing fever, weight loss, and anemia in up to 20-40% of cases pre-ART.[101]Noninfectious complications encompass AIDS-defining malignancies and organ dysfunctions linked to chronic immune dysregulation. Kaposi's sarcoma, driven by human herpesvirus 8 (HHV-8), forms multifocal vascular lesions in skin, mucosa, viscera, and lungs, historically comprising 50% of initial AIDS diagnoses in the U.S. but now rarer with ART.[106]Non-Hodgkin lymphoma, often EBV-associated aggressive B-cell types, infiltrates CNS or systemic sites at elevated rates (up to 100-fold increase in HIV), presenting with rapid masses, B symptoms, and poor prognosis without intervention.[107]Invasive cervical cancer, tied to human papillomavirus (HPV) persistence, qualifies as AIDS-defining and shows higher incidence in HIV-positive women due to impaired viral clearance.[108]Neurological complications include HIV-associated neurocognitive disorder (HAND), ranging from asymptomatic impairment to dementia, affecting up to 50% of untreated patients via direct viral effects on the brain or secondary inflammation; symptoms involve memory deficits, motor slowing, and behavioral changes.[109] HIV encephalopathy manifests as progressive subcortical dementia with atrophy on imaging, while peripheral neuropathies cause distal sensory loss and pain in 30-60% of advanced cases.[110] Wasting syndrome, defined by involuntary weight loss >10% plus fever or diarrhea, stems from cytokine dysregulation and malabsorption, historically contributing to 20-30% of AIDS mortality.[111] These entities underscore the multisystemic impact of untreated HIV, with immune reconstitution under ART sometimes provoking inflammatory syndromes exacerbating presentations.[112]
Treatment
Antiretroviral Therapy
Antiretroviral therapy (ART) consists of combinations of medications that target multiple stages of the HIV replication cycle, including reverse transcription, integration, protease activity, and entry, thereby suppressing viral load and preserving immune function.[4] Introduced in the late 1980s, ART has evolved from monotherapy to multidrug regimens, reducing HIV-related mortality by over 50% in treated populations since the mid-1990s.[113] The first antiretroviral drug, zidovudine (AZT), was approved by the FDA on March 19, 1987, for advanced AIDS cases, demonstrating initial reductions in mortality but limited efficacy due to rapid viral resistance.[114]The advent of highly active antiretroviral therapy (HAART) in 1996 marked a pivotal shift, combining at least three drugs from two classes—typically two nucleoside reverse transcriptase inhibitors (NRTIs) with a protease inhibitor (PI) or non-nucleoside reverse transcriptase inhibitor (NNRTI)—to achieve sustained viral suppression.[115] This approach restored CD4 T-cell counts and averted progression to AIDS in adherent patients, transforming HIV into a manageable chronic condition rather than a fatal illness.[116] By 2015, ART access had prevented an estimated 9.5 million deaths globally, with economic returns of $3.50 per dollar invested through reduced healthcare costs and productivity gains.[113]Current U.S. Department of Health and Human Services guidelines, updated September 25, 2025, recommend initiating ART immediately upon HIV diagnosis for all individuals, regardless of CD4 count or viral load, to maximize immune recovery and prevent transmission.[117] Preferred initial regimens include integrase strand transfer inhibitor (INSTI)-based single-tablet combinations such as bictegravir/emtricitabine/tenofovir alafenamide (BIC/FTC/TAF) or dolutegravir plus lamivudine (DTG/3TC), administered once daily for optimal adherence.[118] These regimens achieve undetectable viral loads (<50 copies/mL) in over 90% of adherent patients within six months, correlating with "undetectable equals untransmittable" (U=U) outcomes in serodiscordant couples.[119] WHO guidelines, updated September 12, 2025, similarly endorse dolutegravir-based first-line therapy for adults, emphasizing simplified dosing to support long-term retention in resource-limited settings.[120]Virologic monitoring every 4-6 months is essential to confirm suppression and detect resistance, which arises from HIV's high mutation rate (approximately 3 × 10^-5 per nucleotide per replication cycle) and incomplete adherence.[121] Resistance mutations, such as M184V in reverse transcriptase or major PI substitutions, render specific drugs ineffective, necessitating genotypic testing and regimen switches to second-line options like boosted darunavir with an NRTI backbone.[122] Globally, pretreatment resistance to NNRTIs affects up to 10% of new diagnoses in high-burden areas, underscoring the need for INSTI prioritization.[11]Adverse effects vary by drug class: NRTIs like tenofovir can cause renal toxicity and bone density loss; PIs and older NNRTIs are linked to dyslipidemia, insulin resistance, and hepatotoxicity; while modern INSTIs generally offer better tolerability with rare neuropsychiatric effects like insomnia or weight gain.[123] Long-term risks include accelerated aging-related comorbidities, such as cardiovascular disease, mitigated by lifestyle interventions and regimen optimization.[124] Lifelong adherence, exceeding 95% for efficacy, remains critical, as interruptions foster resistance and rebound viremia.[4]
Prophylaxis and Symptom Management
Primary prophylaxis against opportunistic infections (OIs) targets HIV-infected individuals with severely impaired immunity, particularly CD4 counts below specific thresholds, to prevent initial episodes despite antiretroviral therapy (ART). The U.S. Department of Health and Human Services (DHHS) and National Institutes of Health (NIH) guidelines, updated October 8, 2024, recommend trimethoprim-sulfamethoxazole (TMP-SMX) as first-line for Pneumocystis jirovecii pneumonia (PCP) when CD4 is <200 cells/µL, administered as one double-strength tablet daily; alternatives include dapsone, atovaquone, or aerosolized pentamidine for those intolerant to TMP-SMX.[125] For toxoplasmic encephalitis, TMP-SMX provides dual prophylaxis when CD4 is <100 cells/µL and Toxoplasma-specific IgG is positive. Mycobacterium avium complex (MAC) prophylaxis uses azithromycin 1200 mg weekly or clarithromycin 500 mg twice daily if CD4 is <50 cells/µL. Prophylaxis is discontinued once ART achieves immune reconstitution, such as CD4 >200 cells/µL for PCP for at least 3 months, reducing pill burden and resistance risk.[125]
Secondary prophylaxis follows acute OI treatment to prevent relapse, continuing until sustained CD4 recovery; for example, lifelong suppressive therapy for cryptococcal meningitis uses fluconazole 200 mg daily after induction. Vaccination strategies complement prophylaxis: inactivated vaccines like pneumococcal (PCV20 or PPSV23) and influenza are recommended annually, while live vaccines (e.g., yellow fever) are contraindicated if CD4 <200 cells/µL.[126]Symptom management in HIV focuses on alleviating ART-related adverse effects, HIV-driven constitutional symptoms, and complications from comorbidities, emphasizing multidisciplinary supportive care to maintain quality of life. Common manifestations include peripheral neuropathy (affecting up to 30% of patients on certain ART regimens like stavudine historically, though less with modern integrase inhibitors), managed with tricyclic antidepressants such as amitriptyline 25-50 mg nightly or anticonvulsants like gabapentin titrated to 1800-3600 mg daily, alongside ART optimization.[127] HIV-associated wasting, characterized by unintentional weight loss >10% body weight, responds to nutritional supplementation, resistance exercise, and appetite stimulants like megestrol acetate 400-800 mg daily, though evidence shows modest gains of 1-2 kg without reversing mortality risk if CD4 remains low.[127] Gastrointestinal symptoms such as chronic diarrhea from opportunistic pathogens or ART (e.g., protease inhibitors) are addressed by pathogen-specific therapy, antidiarrheals like loperamide, or switching regimens; persistent cases warrant endoscopy to exclude malignancy.[125]In advanced disease, palliative approaches integrate pain control (WHO ladder: non-opioids first, opioids like morphine for severe pain, with adjuncts for neuropathic components) and psychosocial support, as outlined in guidelines prioritizing symptom relief over curative intent when ART fails.[128] Fatigue and cognitive issues, prevalent in 50-70% of untreated or virologically suppressed patients, benefit from sleep hygiene, exercise, and ruling out anemia or depression via targeted screening, though no single pharmacologic agent is universally effective.[126] Adherence counseling and monitoring mitigate symptoms from non-adherence, which correlates with 2-3 fold higher OI incidence.[129] Overall, effective management hinges on viral suppression via ART, with prophylaxis and supportive measures bridging gaps in immune recovery.[130]
Recent Therapeutic Advances
In recent years, long-acting injectable antiretroviral therapies (LA-ART) have emerged as a significant advancement for HIV treatment, addressing adherence challenges associated with daily oral regimens. Cabotegravir and rilpivirine, approved by the FDA in 2021 for adults and extended to adolescents weighing at least 35 kg in 2022, allow for monthly or bimonthly intramuscular injections after an oral lead-in phase, achieving viral suppression comparable to oral therapy in phase 3 trials like FLAIR and ATLAS.[131][132] These formulations maintain sustained plasma levels, with 48-week viral suppression rates exceeding 90% in virologically suppressed patients switching from oral ART.[133] Further developments include ViiV Healthcare's investigational ultra-long-acting cabotegravir variant, enabling dosing three times annually, presented in late 2024 as part of efforts to extend intervals beyond current standards.[134]Capsid inhibitors represent a novel class of antiretrovirals with prolonged activity. Lenacapavir (Sunlenca), approved by the FDA in December 2022 for heavily treatment-experienced adults with multidrug-resistant HIV, disrupts viral capsid assembly and nuclear import, demonstrating superior efficacy over optimized background regimens in the CAPELLA trial, where 83% achieved undetectable viral loads by week 52.[135] Its subcutaneous administration every six months positions it for potential maintenance therapy, with ongoing phase 3 data at CROI 2025 exploring twice-yearly regimens in broader populations.[135] Gilead's pipeline includes combinations integrating lenacapavir with other agents for simplified, long-duration treatment options.[136]Broadly neutralizing antibodies (bNAbs) offer promise for simplified or latency-targeted therapy. A 2024 phase 1b trial published in Nature Medicine tested a triple bNAb combination (3BNC117, 10-1074, and 10E8V4), achieving rapid viral suppression in viremic participants when combined with ART, with broader neutralization coverage reducing escape risk compared to single agents.[137] In humanized mouse models and early human studies, bNAbs like these have delayed viral rebound post-ART interruption for months, supporting their role in functional cure strategies.[138]ViiV Healthcare and others are advancing bNAb formulations for long-acting delivery, potentially enabling treatment-free periods.[139]Gene-editing approaches toward curative intent have progressed in clinical testing. EBT-101, an AAV-delivered zinc finger nuclease targeting CCR5 and HIV proviral DNA, received FDA fast-track designation in July 2023, with phase 1 trials assessing safety and reservoir reduction ongoing through March 2025.[140] While stem cell transplants with CCR5-delta32 mutations have cured at least 10 individuals by mid-2025, primarily in cancer patients, these remain non-scalable; bNAb-assisted "shock-and-kill" strategies continue in preclinical refinement to activate and eliminate latent reservoirs.[141][142] These advances prioritize empirical efficacy data over speculative outcomes, with trials emphasizing measurable viral control metrics amid persistent challenges like resistance and access disparities.[143]
History
Evolutionary Origins
HIV-1 and HIV-2 arose independently through multiple cross-species transmissions of simianimmunodeficiency viruses (SIVs) from non-human African primates to humans, likely facilitated by bushmeat hunting practices involving blood exposure.[5] Phylogenetic analyses of viral genomes confirm that these zoonotic events represent the primary origins, with no evidence supporting alternative non-zoonotic hypotheses in the empirical record.[144] The viruses adapted to human hosts via genetic recombinations and mutations, enabling sustained transmission and eventual pandemics.[145]The predominant HIV-1 strains, particularly group M responsible for the global pandemic, originated from SIVcpz infecting chimpanzees (Pan troglodytes troglodytes) in southeastern Cameroon.[5] Molecular clock dating estimates the initial transmission to humans occurred between 1890 and 1920, with subsequent diversification in the Kinshasa region of the Democratic Republic of Congo by the 1920s.[146] The earliest direct evidence of HIV-1 in a human sample dates to 1959 from a preserved blood specimen in Kinshasa.[147] HIV-1 groups N, O, and P also trace to chimpanzee or closely related gorilla SIV lineages, though group M's adaptations—such as enhanced transmissibility—drove widespread establishment.[6][148]HIV-2, less virulent and primarily endemic to West Africa, derives from SIVsmm in sooty mangabeys (Cercocebus atys).[149] At least eight independent transmissions occurred during the first half of the 20th century, with subtypes A and B emerging around 1940.[150] Phylogenetic reconstruction places the most recent common ancestor of SIVsmm and HIV-2 in mangabeys circa 1931, reflecting regional human-primate contact patterns.[151] Unlike HIV-1, HIV-2's limited global spread correlates with lower replication efficiency in humans.[152]
Discovery and Early Outbreaks
The first recognized cases of what would later be termed acquired immunodeficiency syndrome (AIDS) emerged in the United States in 1981. On June 5, 1981, the Centers for Disease Control and Prevention (CDC) published a report in its Morbidity and Mortality Weekly Report (MMWR) describing five cases of Pneumocystis carinii pneumonia (PCP), an opportunistic infection typically seen only in severely immunocompromised individuals, among previously healthy homosexual men in Los Angeles. [153] These patients exhibited severe T-cell deficiencies, marking the initial signal of an emerging epidemic. [154] By July 1981, additional reports detailed clusters of Kaposi's sarcoma, a rare cancer, among gay men in New York City and California, further indicating a common underlying immunodeficiency. [155]Throughout 1981 and into 1982, cases expanded beyond initial clusters, with over 150 AIDS diagnoses reported by mid-1982, including among hemophiliacs, Haitian immigrants, and injection drug users, suggesting transmission via blood and sexual contact rather than solely lifestyle factors. [156] The syndrome was formally named AIDS in September 1982 following a CDC case definition that emphasized opportunistic infections and Kaposi's sarcoma in the absence of known immunosuppression causes. [155] Phylogenetic analyses later estimated the U.S. HIV-1 subtype B epidemic's most recent common ancestor around 1969 (95% confidence interval: 1966–1972), implying cryptic circulation for over a decade before recognition. [157] In Africa, where HIV-1 originated from simian immunodeficiency virus cross-species transmission, early human infections dated to the early 20th century, but clinical outbreaks were not systematically reported until 1983, with the first confirmed AIDS case in Rwanda that year. [158]The causative agent, human immunodeficiency virus (HIV), was isolated in 1983 by Luc Montagnier and Françoise Barré-Sinoussi at the Pasteur Institute in Paris. In January 1983, they cultured a novel retrovirus, termed lymphadenopathy-associated virus (LAV), from a lymph node biopsy of a 59-year-old French patient with persistent lymphadenopathy preceding AIDS. [159] This isolation was published on May 20, 1983, in Science, demonstrating the virus's cytopathic effects on T-lymphocytes. [160] Independently, Robert Gallo at the National Cancer Institute reported isolating a cytopathic retrovirus, named HTLV-III, from AIDS patients in May 1984, enabling scalable propagation and serological testing that confirmed HIV as the etiological agent. [161] A subsequent U.S. Department of Health and Human Services announcement on April 23, 1984, credited Gallo's work with identifying the virus's cause of AIDS, though disputes over priority and sample sharing led to Franco-U.S. agreements in 1987 and the 2008 Nobel Prize in Physiology or Medicine awarded solely to Montagnier and Barré-Sinoussi. [162] By late 1984, over 3,000 AIDS cases had been reported in the U.S., with mortality nearing 1,300, underscoring the epidemic's rapid progression. [163]
Public Health Response Milestones
The Centers for Disease Control and Prevention (CDC) first documented cases of the syndrome that became known as AIDS on June 5, 1981, reporting Pneumocystis carinii pneumonia in five previously healthy young homosexual men in Los Angeles, which prompted initial surveillance efforts focused on clusters among men who have sex with men.[153] By July 1981, additional CDC reports identified Kaposi's sarcoma in 26 gay men across New York and California, expanding case tracking and highlighting immune deficiency patterns.[155] In September 1982, the CDC formalized a case definition for acquired immune deficiency syndrome (AIDS) to standardize reporting, enabling national surveillance that by December 1982 had recorded 593 cases and 243 deaths in the United States.[158]The approval of the first enzyme-linked immunosorbent assay (ELISA) for HIV antibodies by the Food and Drug Administration (FDA) on March 2, 1985, marked a pivotal advance in screening, leading to mandatory testing of blood donations and reducing transfusion-related transmissions from 10% of cases pre-1985 to near zero thereafter.[155] In May 1987, the World Health Organization (WHO) launched the Global Programme on AIDS (GPA), coordinating international surveillance, prevention, and education across 107 countries with an initial budget of $90 million, which grew to address heterosexual transmission in developing regions.[164]Domestic U.S. policy advanced with the Ryan White Comprehensive AIDS Resources Emergency (CARE) Act, signed into law on August 18, 1990, allocating federal funds—initially $225 million—for care, support services, and epidemiology in underserved areas, serving over 500,000 people by 2000.[155] The establishment of the Joint United Nations Programme on HIV/AIDS (UNAIDS) in January 1996 unified six UN agencies to streamline global responses, setting targets like universal access to prevention and treatment, which by 2015 expanded antiretroviral therapy to 15 million people worldwide.[158]In 2003, U.S. President George W. Bush announced the President's Emergency Plan for AIDS Relief (PEPFAR), authorizing $15 billion over five years for antiretroviral treatment, prevention of mother-to-child transmission, and care in 15 focus countries, primarily in sub-Saharan Africa, where it supported over 1.2 million on therapy by 2008 and averted an estimated 1.1 million pediatric infections through 2013.[158] The UN General Assembly's 2011 Political Declaration on HIV/AIDS committed to reducing new infections by 50% by 2015, influencing national programs that integrated HIV testing into routine health services and scaled up pre-exposure prophylaxis (PrEP) approvals, such as the FDA's 2012 endorsement of Truvada for high-risk individuals.[155]
Controversies
Denialism and Alternative Theories
HIV denialism refers to the rejection of the established causal link between human immunodeficiency virus (HIV) and acquired immunodeficiency syndrome (AIDS), positing instead that HIV is a harmless passenger virus or that AIDS results from factors such as recreational drug use, malnutrition, or antiretroviral medications themselves.[165] Proponents often argue that HIV fails to satisfy Koch's postulates for causation, citing low viral loads in infected individuals and the absence of direct T-cell killing by the virus in vitro.[166] These views emerged in the late 1980s amid early uncertainties in HIV research but have persisted as fringe positions despite accumulating contradictory evidence.[167]A primary architect of denialist arguments is molecular biologist Peter Duesberg, who in 1987 contended that American and European AIDS cases stem from long-term recreational drug consumption—particularly nitrite inhalants—and the toxicity of anti-HIV drugs like AZT, rather than HIV infection.[168] Duesberg maintained that HIV's isolation and serological detection merely correlate with high-risk behaviors, not disease causation, and predicted that AIDS would not occur in non-drug-using populations, a claim later falsified by cases in hemophiliacs and transfusion recipients without such histories.[166] His hypothesis gained limited traction in academic circles initially but faced professional repercussions, including loss of funding, as it diverged from empirical data showing HIV's role in immune depletion.[169]Other figures include biochemist Kary Mullis, Nobel laureate for inventing PCR, who expressed skepticism that HIV directly causes AIDS and criticized viral load tests as unreliable for diagnosis due to PCR's sensitivity amplifying non-pathogenic fragments.[170] Mullis viewed HIV as a marker for underlying conditions rather than the etiologic agent.[171] Politically, South African President Thabo Mbeki amplified denialism from 1999 to 2005 by questioning the HIV-AIDS paradigm in international forums, delaying antiretroviral rollout and favoring unproven nutritional interventions like vitamins and African remedies.[172] This policy contributed to an estimated 330,000 preventable deaths and 35,000 excess mother-to-child transmissions between 2000 and 2005, as government obstruction hindered access to proven therapies.[173]Alternative theories extend to claims that antiretroviral therapy itself induces AIDS-like symptoms through mitochondrial toxicity or immune suppression, or that oxidative stress from lifestyle factors explains immune collapse independently of HIV.[174] These positions have influenced small activist groups and online communities, sometimes intersecting with conspiracy narratives alleging pharmaceutical profiteering or fabricated epidemics.[175]The scientific consensus, grounded in longitudinal cohort studies, rejects denialism, affirming HIV as the necessary cause of AIDS through mechanisms like CD4+ T-cell depletion via viral replication and chronic immune activation.[176] Key evidence includes the restoration of immune function and prevention of progression to AIDS upon viral suppression with antiretrovirals, as seen in randomized trials reducing mortality by over 80% since 1996; epidemiological patterns matching HIV transmission routes rather than solely drug use; and primate models where SIV (HIV's simian analog) induces AIDS-like pathology absent confounding factors.[171] Denialist predictions, such as AIDS vanishing without HIV treatment, failed as untreated infection consistently progresses to opportunistic infections and death, while prophylaxis against confounders like drugs does not avert decline in HIV-positive individuals.[166] Mainstream institutions, despite potential biases in narrative alignment, align on this causation based on replicable molecular, clinical, and interventional data spanning decades.[2]
Origins Hypotheses Beyond Zoonosis
The oral polio vaccine (OPV) hypothesis, proposed by journalist Edward Hooper in his 1999 book The River, posits that HIV-1 originated from experimental OPV trials conducted in the Belgian Congo (now Democratic Republic of Congo) and Rwanda between 1957 and 1960, where vaccines allegedly grown in chimpanzee kidney cells contaminated with simian immunodeficiency virus (SIVcpz) were administered to over a million people.[177] Hooper argued that the use of chimpanzee tissues from SIV-endemic regions like Stanleyville, combined with evidence of tissue pooling and inadequate sterilization, facilitated cross-species transmission, with phylogenetic clustering of early HIV strains near trial sites supporting temporal and geographic alignment.[178] He cited archival records indicating Hilary Koprowski's team sourced kidneys from local chimps, potentially introducing SIV variants that mutated into HIV-1 group M, the pandemic strain, rather than relying solely on rare bushmeat hunting events.[179]Critiques of the OPV hypothesis emphasize phylogenetic dating analyses placing the most recent common ancestor (MRCA) of HIV-1 group M in the early 20th century, around 1908–1933, predating the OPV campaigns by decades and aligning with colonial-era disruptions like railway construction that increased bushmeat exposure.[180] Hooper contested these molecular clock estimates as flawed due to assumptions of constant mutation rates and sampling biases, advocating for re-evaluation based on epidemiological patterns in the Democratic Republic of Congo.[180] Independent testing of archived vaccine samples in 2000–2001 found no detectable chimp DNA or SIV traces, and production records confirmed most CHAT vaccine lots used macaque monkey kidneys, not chimpanzee, with any chimp use limited and non-endemic.[181] Subsequent serological surveys and genetic studies reinforced natural zoonotic spillover from chimpanzees via cutaneous exposure during hunting, without requiring iatrogenic amplification, though proponents like Hooper maintain that undetected contamination remains possible given incomplete archival access.[5]Man-made bioweapon theories, circulating since the 1980s, claim HIV was engineered in a U.S. militarylaboratory, often linked to Fort Detrick or hepatitis B vaccine trials, as a deliberate pathogen targeting specific populations.[182] These narratives gained traction through disinformation campaigns, including a 1986 East German Stasi operation via Indian journalist Rakesh Chandra (code-named "HIV"), which fabricated evidence of U.S. creation to undermine Western credibility during the Cold War.[183] Proponents cited the virus's rapid global spread and perceived genetic anomalies, such as its glycoproteinenvelope, as hallmarks of laboratory design, with surveys in the 1990s showing up to 25–30% belief in government lab origins among certain U.S. demographics.[184]Such theories lack empirical support, as HIV-1's genome exhibits natural recombination patterns from SIVcpz and SIVgor, inconsistent with de novo synthesis, and archival virological records show no pre-1980s evidence of HIV-like constructs in labs.[182]Genetic divergence analyses trace multiple HIV lineages to independent primate-to-human jumps between 1884 and 1924, predating modern recombinant DNA technology (developed in the 1970s), and epidemiological modeling attributes pandemic scale to 20th-century urbanization and medical practices amplifying low-prevalence zoonoses, not artificial release.[185] These hypotheses persist in fringe discourse but are contradicted by cross-verified phylogenetic trees and lack primary documentation, often relying on anecdotal or forged claims from state-sponsored propaganda.[183]
Critiques of Prevention Policies
Critiques of HIV prevention policies have centered on their overreliance on condom distribution and biomedical interventions like pre-exposure prophylaxis (PrEP), which some analysts argue fail to address root behavioral drivers of transmission and may inadvertently encourage riskier sexual practices. In sub-Saharan Africa, where HIV prevalence peaked in the 1990s, Uganda's ABC strategy—emphasizing abstinence (A), fidelity in sexual partnerships (B), and condoms (C) as a last resort—correlated with a decline in national HIV prevalence from approximately 18-21% in 1992 to 5% by 2001, primarily through reduced numbers of sexual partners and delayed sexual debut rather than widespread condom adoption.[186] This contrasts with policies in other high-prevalence countries, where condom-focused campaigns yielded limited incidence reductions, prompting arguments that Western-influenced programs undervalued cultural and moral appeals to abstinence and monogamy, potentially due to ideological biases prioritizing sexual liberation over restrictive norms.[187][188]Condom promotion, a cornerstone of global HIV strategies since the 1980s, faces criticism for inconsistent real-world efficacy due to user errors such as breakage, slippage, and improper application, which occur in 1-5% of uses and substantially undermine protection against HIV and other sexually transmitted infections.[189][190] Studies indicate that these mechanical failures, combined with dual protection gaps (e.g., against non-HIV STIs), contribute to persistent transmission in high-risk groups like female sex workers and men who have sex with men, where correct and consistent use remains below 50% despite education campaigns.[191] Critics contend that policies emphasizing free condom distribution often neglect complementary behavioral interventions, leading to stable or rising incidence rates; for instance, U.S. HIV diagnoses among men who have sex with men declined only modestly (18%) from 2015-2019 amid intensified promotion, yet condom use rates hovered around 30-50% in surveys.[192]The rollout of PrEP since 2012 has drawn scrutiny for inducing risk compensation, where users perceive reduced HIV vulnerability and increase condomless sex or partner numbers, resulting in elevated sexually transmitted infection rates—up to 20-30% higher in some cohorts—without proportional HIV prevention gains if adherence lapses.[193][194] Modeling and observational data suggest this moral hazard effect offsets PrEP's biomedical benefits, particularly among men who have sex with men, where STI incidence rose post-adoption despite high efficacy (99% when adherent); undetected acute HIV at initiation can also foster drug-resistant strains, complicating treatment.[195][196] Proponents of behavioral primacy argue that such policies, often advanced by public health institutions with progressive leanings, prioritize access over accountability, sidelining evidence from abstinence-inclusive models like Uganda's that achieved steeper declines through partner reduction.[197] Overall, these critiques highlight a pattern where resource-intensive biomedical and distributive approaches underperform against multifaceted behavioral strategies, with global HIV incidence plateauing at around 1.5 million new cases annually as of 2022 despite $20+ billion in yearly prevention funding.30035-2/fulltext)
Research Frontiers
Vaccine Efforts
Development of an HIV vaccine has proven exceptionally challenging due to the virus's high mutation rate, which generates extensive genetic variability exceeding that of any other virus, complicating the targeting of conserved epitopes.[198] HIV integrates into host DNA as a provirus, evading traditional immune clearance mechanisms and limiting vaccine platforms like live attenuated viruses that succeed against other pathogens.[199] Additionally, HIV primarily infects CD4+ T cells, the coordinators of adaptive immunity, impairing the host's ability to mount sterilizing immunity, while natural infection rarely confers protection, with only a small fraction of individuals achieving elite control through rare broadly neutralizing antibodies (bNAbs).[200] These factors have resulted in over four decades of research yielding no licensed preventive vaccine, despite substantial investment.[201]Early efforts from 1987 focused on subunit vaccines targeting the envelope glycoprotein gp160 or gp120 to elicit neutralizing antibodies, but phase I trials showed no significant efficacy in preventing infection.[202] By the 1990s and 2000s, trials shifted toward cellular immunity with vectors like canarypox (ALVAC) combined with gp120 boosts, culminating in the RV144 trial in Thailand (2003–2009), which enrolled 16,402 high-risk adults and reported a modest 31.2% efficacy against HIV acquisition after 3.5 years, the first positive signal in human trials without increasing infection risk.[198] Follow-up trials, such as HVTN 702 in South Africa (2016–2020) testing an improved ALVAC-gp120 regimen, failed to replicate efficacy, halted early due to futility with no reduction in infections among 5,400 participants.[203] Similarly, adenovirus-5 vectored vaccines like STEP and Phambili (2007) were terminated after interim analyses showed either no benefit or increased risk in some subgroups, highlighting potential immune interference from pre-existing vector immunity.[204]Contemporary strategies emphasize inducing bNAbs, which neutralize diverse HIV strains by targeting conserved sites on the envelope trimer, but require sequential immunization to guide rare B-cell precursors through affinity maturation—a process unachievable in standard vaccines.[200] Germline-targeting approaches, using engineered immunogens to activate naive B cells, advanced in 2025 with IAVI-led phase 1 trials demonstrating proof-of-concept for stepwise regimens that elicited bnAb precursors in humans, published in Science on May 15.[205] mRNA platforms, adapted from COVID-19 successes, showed potent CD4+ T-cell and antibody responses in early 2025 trials, overcoming envelope instability challenges by stabilizing trimers, though long-term breadth against variants remains unproven.[206][207] A phase 1 trial launched in Africa on August 4, 2025, evaluates novel candidates in 120 adults, including those with HIV, prioritizing safety and immunogenicity in high-prevalence settings.[208] Despite these advances, experts note persistent hurdles, including the need for animal models better mimicking humantransmission and the field's vulnerability to funding cuts, as seen in U.S. NIAID's 2025 cancellation of key consortia.[209] No trial has yet achieved durable, high-efficacy protection, underscoring the requirement for multifaceted immune responses beyond antibodies alone.[210]
Cure and Eradication Strategies
HIV persists in latent reservoirs within resting CD4+ T cells and tissues such as lymphoid organs and the central nervous system, where integrated proviral DNA evades immune detection and antiretroviral therapy (ART), rendering complete eradication challenging despite viral suppression to undetectable levels.[211][212] These reservoirs decay slowly, with half-lives estimated at years to decades, complicating efforts to achieve sustained remission without ART.[213] Tissue-specific persistence, including in gut-associated lymphoid tissue and brain macrophages, further hinders uniform targeting, as quantification and characterization of reservoirs remain imprecise due to methodological limitations.[214][215]Stem cell transplantation has demonstrated proof-of-concept cures in rare cases, primarily when donor cells carry the homozygous CCR5-Δ32 mutation conferring HIV resistance, though successes without this mutation suggest broader mechanisms like graft-versus-virus effects.[216] As of August 2025, approximately 10 individuals worldwide, including the Berlin patient (cured in 2008 after 2006 transplants for leukemia), London patient (2019), Geneva patient (remission announced 2024 without CCR5 mutation), and a seventh German case (2024), have achieved long-term remission post-transplant, verified by analytical treatment interruptions showing no viral rebound.[141][217] However, this approach is not scalable for the 39 million people living with HIV due to high mortality risks (up to 10-20% from transplant-related complications), donor scarcity, and unsuitability for most patients without concurrent malignancies.[218]Latency-reversing agents (LRAs) aim to reactivate proviruses in the "shock and kill" strategy, exposing infected cells to immune clearance or ART, but clinical trials have shown only modest reservoir reductions without achieving cure.[219] Agents like histone deacetylase inhibitors (e.g., vorinostat) and protein kinase C agonists have induced HIV transcription in vivo but failed to clear reservoirs effectively, partly due to incomplete reactivation, T-cell exhaustion, and lack of immune enhancement. Ongoing trials, such as those combining LRAs with immune modulators (e.g., NCT06240520 initiated 2024), explore optimized combinations, yet challenges persist in pharmacovigilance and avoiding non-specific T-cell activation.[220]Gene-editing technologies, particularly CRISPR-Cas9, target HIV proviral DNA excision or host factors like CCR5 to prevent infection. In nonhuman primates, a single CRISPR injection eliminated SIV genomes from most tissues in 2023 models, reducing viral loads post-ART interruption.[221] Human trials, including EBT-101 (FDA fast-track 2023), demonstrated safety and targeted DNA cleavage in blood by mid-2024 but did not prevent rebound upon ART cessation in early participants, highlighting needs to address off-target effects, delivery to reservoirs, and viral escape. Dual-guide CRISPR approaches show promise in inhibiting replication without driving resistance, as confirmed in 2025 rebound studies.[222]Broadly neutralizing antibodies (bNAbs) neutralize diverse HIV strains and delay rebound when combined with ART, potentially shrinking reservoirs via antibody-dependent cellular cytotoxicity. A 2024 trial of triple bNAb combinations suppressed viremia in 70-90% of participants during analytical interruptions but required ongoing dosing to prevent escape mutants.[137] Over 150 registered trials by 2025 evaluate bNAbs for cure adjuncts, with next-generation variants enhancing coverage against global clades.[138][223]Emerging strategies exploit HIV's biology, such as inducing cellular intolerance to reactivation-induced damage to trigger apoptosis of latent cells, as shown in October 2025 preclinical models. Despite advances, no scalable cure exists; eradication efforts prioritize ART access and prevention, with tissue reservoirs posing ongoing barriers to global elimination.[224][225]
Emerging Resistance and Long-Term Challenges
Despite widespread adoption of antiretroviral therapy (ART), HIV's error-prone reverse transcriptase enzyme generates mutations at a rate of approximately 3 × 10^{-5} per nucleotide per replication cycle, facilitating the emergence of drug-resistant variants.[226] Transmitted drug resistance (TDR) prevalence among newly diagnosed individuals varies globally; in the United States, overall TDR rates reached 18.9% in one large study, with non-nucleoside reverse transcriptase inhibitor (NNRTI) resistance predominant, while another analysis reported 13.5% with an increasing trend over time.[227][228] In low- and middle-income countries, acquired resistance to first-line drugs like efavirenz and nevirapine affects 31-33% of treatment-experienced patients, though integrase strand transfer inhibitors like dolutegravir (DTG) show low resistance rates (under 1% pretreatment in most surveys as of 2024).[229][230] Declines in resistance have occurred in high-resource settings, with U.S. rates dropping nearly 17% from 2018 to 2024 due to optimized regimens emphasizing high-barrier drugs like DTG.[231] However, suboptimal adherence and treatment interruptions—reported in up to 20-30% of patients annually—accelerate resistance selection, particularly in resource-limited areas where monitoring is inconsistent.[232][11]Long-term management demands lifelong ART adherence, yet challenges persist, including pill fatigue, with adherence rates falling below 95% (the threshold for sustained suppression) in 10-20% of patients over years.[233][234] Comorbidities exacerbate issues; HIV-accelerated aging increases risks of cardiovascular disease, neurocognitive impairment, and malignancies, even with viral suppression, due to persistent immune activation and inflammation.[235] Drug toxicities, such as renal or bone effects from tenofovir, affect 5-10% long-term, necessitating regimen switches.[236] Access barriers in low-income regions hinder control, with only 77% of people living with HIV on ART as of 2023, fueling ongoing transmission of resistant strains.[237]The latent viral reservoir—primarily in resting CD4+ T cells—poses the greatest eradication barrier, decaying with a half-life of about 44 months under ART, implying over 70 years for natural clearance in most patients.[238] These reservoirs, established early in infection, harbor replication-competent virus shielded from ART and immune detection, with clonal expansion sustaining persistence despite suppression.[239] "Shock and kill" strategies to reactivate and eliminate latently infected cells have failed in trials due to incomplete activation, off-target effects, and reservoir rebound post-treatment interruption.[211] Emerging approaches like long-acting injectables (e.g., cabotegravir/rilpiviril) aim to improve adherence but do not address latency, underscoring the need for reservoir-targeting interventions amid rising treatment complexity.[234][240]