From Ancient to Emerging Infections: The Odyssey of Viruses in the Male Genital Tract
Abstract
The male genital tract (MGT) is the target of a number of viral infections that can have deleterious consequences at the individual, offspring, and population levels. These consequences include infertility, cancers of male organs, transmission to the embryo/fetal development abnormalities, and sexual dissemination of major viral pathogens such as human immunodeficiency virus (HIV) and hepatitis B virus. Lately, two emerging viruses, Zika and Ebola, have additionally revealed that the human MGT can constitute a reservoir for viruses cleared from peripheral circulation by the immune system, leading to their sexual transmission by cured men. This represents a concern for future epidemics and further underlines the need for a better understanding of the interplay between viruses and the MGT. We review here how viruses, from ancient viruses that integrated the germline during evolution through old viruses (e.g., papillomaviruses originating from Neanderthals) and more modern sexually transmitted infections (e.g., simian zoonotic HIV) to emerging viruses (e.g., Ebola and Zika) take advantage of genital tract colonization for horizontal dissemination, viral persistence, vertical transmission, and endogenization. The MGT immune responses to viruses and the impact of these infections are discussed. We summarize the latest data regarding the sources of viruses in semen and the complex role of this body fluid in sexual transmission. Finally, we introduce key animal findings that are relevant for our understanding of viral infection and persistence in the human MGT and suggest future research directions.
Over 30 contemporary and emerging viruses are released in semen. Both genital (e.g., herpes simplex virus, human papillomavirus) and systemic viruses [e.g., human immunodeficiency virus, hepatitis B virus, Zika virus (ZIKV), Ebola virus (EBOV)] can infect not only the testis, but multiple male genital organs, which can represent viral reservoirs. Viral infections in the male genital tract are frequently silent, and seminal shedding can be prolonged or intermittent, making it difficult to detect.
Viral infections can affect male reproductive function through endocrine disturbances; inflammation; oxidative stress; high fever; or direct testis, epididymis, and accessory gland dysfunctions caused by viral replication.
Clinicians should be aware that a number of acute emerging viruses, including arthropod-borne and life-threatening viruses, can persist in semen despite systemic clearance (e.g., yellow fever virus, Nipah virus) and lead to sexual transmission for extended durations (e.g., ZIKV, EBOV, West-Nile virus).
Screening for viruses in semen and seminal fractions is key to establish potential for transmission to partner and embryo (through determination of infectiousness) and pattern of excretion (through longitudinal samples analysis).
Clinicians should perform this investigation in symptomatic patients, or asymptomatic patients wishing to conceive, according to the epidemiological risk/geographical area and must give appropriate counseling to the patients to avoid infection complications for the patient, his partner, or the future embryo/infant.
I. INTRODUCTION
The male genital tract (MGT) is composed of a series of organs and ducts that ensure male gametes production, storage, and transport. It is also endowed with endocrine functions necessary for the maintenance of the male body phenotype. Alteration of male organ integrity and functions following viral attack can therefore affect both general and reproductive health. The search for a viral etiology to male infertility has expanded over the last two decades with the increase of medically assisted reproduction procedures. In addition to deleterious effects at the individual level, viral infections of the MGT can impact sexual partners and offspring. While the first identification of viruses in human semen and male organs date from the late 1960s to early 1970s (138), concern about the sexual transmission of viruses via semen peaked with the Acquired Immune Deficiency Syndrome (AIDS) pandemic that began in the 1980s and is still ongoing. Lately, emerging viral diseases have renewed the concern about MGT infection and induced a burst of investigations on the subject. Indeed, the unexpected sexual transmission of Zika and Ebola viruses (ZIKV and EBOV, respectively) by cured men revealed that the MGT constitutes a reservoir for viruses. Back in the early 1970s, two papers in Science (93) and the New England Journal of Medicine (695) indicated that men serve as a reservoir for herpes simplex virus (HSV) and cytomegalovirus (CMV). This “reservoir notion” was, however, little developed until the arrival of human immunodeficiency virus (HIV) and the observation of persistence of this virus in the semen of a subset of men with undetectable viremia under effective antiviral therapy (691). However, because all these viruses establish persistent infection in their host, the existence of viral reservoirs is not unexpected, unlike ZIKV and EBOV that do not establish persistent infection in their host. Evidence is currently building up that a number of other emerging and neglected viruses also have the ability to persist in the MGT. Yet the source of viruses in semen and the nature of the MGT reservoirs are unknown.
Investigations on viral infections of the MGT are at the interface between different fields such as reproductive biology/urology and infectious diseases/virology. In this review, we aimed to cover these diverse but interconnected aspects to highlight research needs and to prompt integrated researches.
After a brief overview on viruses (sect. II) and a summary of MGT key features (sect. III), section IV details the routes of entry of viruses into the human MGT and compiles the data available on their target organs and cells at this level. Section V outlines the immune specificities of these organs as well as their equipment and responses against viral agents or mimicry in animal models and humans. The viruses that infect the MGT and their consequences are reviewed in section VI. We begin with the ancient infections of the germline in our ancestors, for which we postulate the implication of male germ cells infection. We summarize the data associating viral infections with human male infertility, as well as studies that investigated a viral etiology for cancers of the male organs, and discuss vertical transmission of viruses through semen. Then we outline evidence for the existence of viral reservoirs in the human MGT. Section VII focuses on semen as a vector of viral dissemination and on the latest data regarding the sources of viruses in this body fluid. Finally, section VIII presents a selection of findings in animals that might enlighten our understanding of viral infections of the human MGT and help us better prepare for upcoming epidemics. In the conclusion, we highlight research needs and propose questions for future research.
II. BRIEF OVERVIEW OF VIRUSES: WHAT ARE THEY? HOW DO THEY INFECT AND SPREAD?
Viruses are obligate intracellular organisms with a simple structure consisting of a nucleic acid (DNA or RNA; single or double stranded; segmented or not; positive-, negative-, or ambi-sense) contained in a proteic shell, the capsid. An outer lipidic envelope derived from host cell membranes is also present in more complex viruses and includes viral glycoproteins involved in the attachment and entry into the host cell. Based on their type of viral genome and mode of replication, viruses are classified (suffix in brackets) by order (-virales), family (-viridae), subfamily (-virinae), genus (-virus), and species [often take the form of (disease) virus] (TABLE 1). Three to four new species are found every year (743), and the taxonomy of viruses is regularly updated by the International Committee on Taxonomy of Viruses (https://talk.ictvonline.org/).
Virus | Family | Genus | Genome | Main Clinical Syndromes |
---|---|---|---|---|
Human papillomavirus (HPV) | Papillomaviridae | α-, β-, γ-Papillomavirus | dsDNA | Warts, precancerous lesions associated with genital, anal, and oropharyngeal cancers |
Herpes simplex virus-1 (HSV-1) | Herpesviridae | Simplexvirus | dsDNA | Herpes labialis, genital herpes |
Herpes simplex virus-2 (HSV-2) | Herpesviridae | Simplexvirus | dsDNA | Genital herpes |
Varicella zoster virus (VZV) | Herpesviridae | Varicellovirus | dsDNA | Chickenpox |
Epstein-Barr virus (EBV) | Herpesviridae | Lymphocryptovirus | dsDNA | Infectious mononucleosis, Burkitt lymphoma, nasopharyngeal carcinoma, and posttransplant lymphoproliferative disease |
Cytomegalovirus (CMV) | Herpesviridae | Cytomegalovirus | dsDNA | Congenital CMV can be life-threatening for immunocompromised patients |
Human herpesvirus-6 (HHV-6) | Herpesviridae | Roseolovirus | dsDNA | Common childhood disease exanthema subitum |
Human herpesvirus-7 (HHV-7) | Herpesviridae | Roseolovirus | dsDNA | Common childhood disease exanthema subitum |
Kaposi sarcoma-associated herpesvirus (KSHV) | Herpesviridae | Rhadinovirus | dsDNA | Kaposi sarcoma, primary effusion lymphoma, Castleman's disease |
Molluscum contagiosum virus (MCV) | Poxviridae | Molluscipoxvirus | dsDNA | Self-limiting papules in genitals and other parts of the body |
Adenovirus (AdV) | Adenoviridae | Mastadenovirus | dsDNA | Conjunctivitis, respiratory or gastroenteric syndromes, meningoencephalitis, urinary tract infections |
Parvovirus B19 | Parvoviridae | Erythrovirus | ssDNA | Fifth disease or erythema infectiosum, hydrops fetalis |
Adeno-associated virus (AAV) | Parvoviridae | Dependovirus | ssDNA | Not known to cause disease. Used in virus-vectored gene-therapy trials |
JC polyomavirus (JCPyV) | Polyomaviridae | Betapolyomavirus | dsDNA | Almost exclusively in immunosuppressed individuals: progressive multifocal leukoencephalopathy |
BK polyomavirus (BKPyV) | Polyomaviridae | Betapolyomavirus | dsDNA | Almost exclusively in immunosuppressed individuals: BK virus associated nephropathy, hemorrhagic cystitis |
Simian virus 40 (SV40) | Polyomaviridae | Betapolyomavirus | dsDNA | Controversy over SV40 involvement in human tumorigenesis |
Merkel cell polyomavirus (MCPyV) | Polyomaviridae | Alphapolyomavirus | dsDNA | Suspected to cause Merkel cell carcinoma (skin cancer) |
Torque teno virus (TTV) | Anelloviridae | Alphatorquevirus | ssDNA | Unknown |
Hepatitis B virus (HBV) | Hepadnaviridae | Orthohepadnavirus | dsDNA (RT) | Hepatitis, cirrhosis, and hepatocellular carcinoma |
Hepatitis D virus (HDV) | Deltavirus | ssDNA | Hepatitis, cirrhosis, and hepatocellular carcinoma upon superinfection and coinfection with HBV | |
Human immunodeficiency virus (HIV) | Retroviridae | Lentivirus | ssRNA (RT) | AIDS |
Human T-lymphotropic virus (HTLV) | Retroviridae | Deltaretrovirus | ssRNA (RT) | Adult T cell leukemia/lymphoma and HTLV-associated myelopathy/tropical spastic paraparesis |
Hepatitis C virus (HCV) | Flaviviridae | Hepacivirus | ssRNA (+) | Hepatitis, cirrhosis, and hepatocellular carcinoma |
Zika virus (ZIKV) | Flaviviridae | Flavivirus | ssRNA (+) | Zika fever, congenital Zika leading to microcephaly and other central nervous system disorders |
West-Nile virus (WNV) | Flaviviridae | Flavivirus | ssRNA (+) | Encephalitis, meningoencephalitis |
Japanese encephalitis virus (JEV) | Flaviviridae | Flavivirus | ssRNA (+) | Encephalitis, meningoencephalitis |
Dengue virus (DENV) | Flaviviridae | Flavivirus | ssRNA (+) | Dengue fever, severe dengue hemorrhagic fever |
Chikungunya virus (CHIKV) | Togaviridae | Alphavirus | ssRNA (+) | CHIKV disease, arthralgia, myalgia |
Coxsackie virus (CoxV B5, A6) | Picornaviridae | Enterovirus | ssRNA (+) | Hand-foot-mouth disease (HFMD), cardiomyopathy, gastrointestinal diseases, central nervous system manifestations |
Hepatitis E virus (HEV) | Hepeviridae | Orthohepevirus A | ssRNA (+) | Hepatitis |
Severe acute respiratory syndrome (SARS) | Coronaviridae | Betacoronavirus | ssRNA (+) | Severe acute respiratory syndrome |
Mumps virus (MuV) | Paramyxoviridae | Rubulavirus | ssRNA (−) | Swelling of the parotid glands, salivary glands, and other epithelial tissues |
Nipah virus (NiV) | Paramyxoviridae | Henipavirus | ssRNA (−) | Acute respiratory illness, fatal encephalitis |
Ebola virus (EBOV) | Filoviridae | Ebolavirus | ssRNA (−) | Hemorrhagic fever |
Marburg virus (MARV) | Filoviridae | Marburgvirus | ssRNA (−) | Hemorrhagic fever |
Andes hantavirus (ANDV) | Hantaviridae | Orthohantavirus | ssRNA (−) | Hantavirus cardiopulmonary syndrome, hantavirus pulmonary syndrome |
Lassa fever virus (LFV) | Arenaviridae | Mammarenavirus | ssRNA (−) ambisense, segmented | Hemorrhagic fever |
Rift Valley fever virus (RVFV) | Bunyaviridae | Phlebovirus | ssRNA (−) ambisense, segmented | A small percentage of infected individuals develops ocular diseases, encephalitis, or hemorrhagic fever |
The life cycle of a selection of viral families is shown in FIGURE 1. The virus host and cell tropism is defined as its capacity to infect specific species and cells. It depends on both viral and cellular characteristics, such as the presence of cognate receptors for the virus attachment molecules, the cellular differentiation state, and expression of antiviral/restriction factors.

FIGURE 1.Life cycle of three representative viral families that infect the male genital tract. A: Flaviviridae (e.g., ZIKV, HCV). B: Retroviridae (e.g., HIV-1/2, HTLV-1/2). C: Herpesviridae (e.g., HSV, CMV). At the host cell level, the main common steps of the life cycle of a virus consists of attachment to the cell surface through receptor recognition, cell entry through either endocytosis/pinocytosis (e.g., Flaviviridae) or a fusion process (e.g., Retroviridae, Herpesviridae), uncoating followed by the release of the nucleic acid in the cytoplasm (Flaviviridae, Retroviriae) or in the nucleus (Herpesviridae), expression and replication of viral genome, assembly of progeny viral particles, and release of newly formed viral particles (virions) through budding or cell lysis. In a cell that is susceptible (i.e., in which the virus may enter through receptor binding), the virus may establish productive infection (cell is fully permissive to viral replication and viral progeny is released, which may lead to cell death resulting in cytolytic infection), restrictive infection (cell is only transiently permissive), latent infection (viral progeny is not produced until active replication is triggered by specific stimuli, as may happen for HSV, VZV, CMV, EBV, and HIV), or abortive infection (replication cannot be completed due to a nonpermissive host or cell, or because the virus is defective). ER, endoplasmic reticulum.
In terms of infection outcome, viral infections are classified as either acute or persistent. In acute infections (e.g., ZIKV), the virus is cleared by the host immune response, while in persistent infections, the virus remains in specific host cells/organs. Persistent infections can be latent [infectious virus not produced between episodes of recurrent disease, e.g., herpes simplex virus (HSV), varicella zoster virus (VZV)], chronic [continued presence of infectious virus following the primary phase, e.g., hepatitis B virus (HBV), HIV] or slow (prolonged incubation period followed by progressive disease). A viral reservoir is a cell type/anatomical site in which replication-competent viruses accumulate and remain over a long period. There are various mechanisms leading to the long-term survival of viruses in the host, but all are based on the ability of the infectious agents to escape from immune system recognition and killing.
Viruses spread among their hosts through horizontal and vertical transmission, through arthropod vectors or fomites, or by cross-species infections (FIGURE 2). Zoonotic infections are those transmitted from animals to humans. More than two-thirds of the viruses that are able to infect humans can also infect animals, and many have a mammalian or avian origin (743).

FIGURE 2.Modes of transmission of viruses. A: inter-host direct transmission can be horizontal (i.e., sexual, parenteral, oro-fecal, respiratory, or through contact with lesions) or vertical, occurring in utero (i.e., gametes-borne or transplacental) or transmitted to the newborn (i.e., at birth or through breastfeeding). B: indirect transmission may be mediated by arthropod vectors (e.g., mosquitoes, ticks, sand flies) or fomites (i.e., inanimate objects able to transmit microbes). C: zoonotic infections are those transmitted from vertebrates to humans, usually through contact with animal fluids, animal bites, or vector transmission. A natural reservoir for a virus is a species in which the pathogen lives and replicates, usually without causing disease.
Emerging infections are defined as infections that have newly appeared in a population, or have existed but are rapidly increasing in incidence or geographic range. Emerging diseases can be caused by previously unknown agents or by agents previously not associated with a specific disease. Re-emerging infections are caused by pathogens whose incidence of disease had significantly declined in the past and has newly appeared in the same or in a different geographical region (349). Several viruses found in the MGT have been termed as emerging [e.g., zoonotic HIV, derived from simian immunodeficiency virus (SIV) in monkeys and that began to spread in the human population in the 1980s] or re-emerging (e.g., recent large outbreaks of ZIKV and EBOV). While viruses constitute only a small portion of the 1,400 human pathogens, they represent more than two-thirds of all new human pathogens (743). Indeed, the ability of viruses to quickly adapt to new hosts makes them the most able to trigger emerging diseases (300).
III. KEY FEATURES OF THE HUMAN MGT
The functional anatomy of the MGT and the main morphological features of the male genital organs and ducts are summarized in FIGURES 3 and 4, with key differences between human and rodent MGT organs highlighted in FIGURE 4C. FIGURE 4 additionally presents the distribution in the human MGT of immune cells, which may have a dual role in viral infections as either targets or defenders.

FIGURE 3.Functional anatomy of the human male genital tract. A: the testis is a highly compartmentalized organ divided into lobules consisting of interstitial tissue and seminiferous tubules, which converge towards the rete testis. Immotile spermatozoa produced in the testis seminiferous tubules during spermatogenesis are expelled through the efferent ducts into the epididymis, where they are matured for 1–21 days as they travel through the epididymis head, body, and tail and are stored. During ejaculation, spermatozoa and epididymis secretions (~10% of seminal plasma) travel through the vas deferens and mix with secretions from the seminal vesicles (~70% of seminal plasma), and then through the ejaculatory ducts which rejoin in the prostate that produces ~30% of seminal plasma. Semen then travels through the bulbourethral gland (Cowper’s gland) and urethra, where the urethral glands (Littré glands) produce the pre-ejaculatory fluid, and exit through the penis meatus. B: semen composition. The box recapitulates the composition of semen, which comprises spermatozoa, seminal leukocytes, exfoliated epithelial cells, and immature germ cells, all bathing in seminal plasma. This dynamic fluid helps the transport and survival of spermatozoa, preserves their fertilizing abilities, and primes an adaptive immune response in the female reproductive tract. Seminal plasma has been shown to directly and indirectly interact with viruses (see sect. VIID).

FIGURE 4.Morphological specificities of human male genital tract (MGT) organs and immune cell distribution. A: histological sections. B: schematics. C: main morphological differences between human and rodents. The interstitial tissue of the testis primarily contains Leydig cells, along with blood and lymphatics vessels (178), immune cells, telocytes (428), fibroblasts, and nerve cells. Leydig cells produce testosterone under the control of the pituitary luteinizing hormone (LH) and also play important paracrine roles for testis functions (508). The seminiferous tubules contain the germ cell-nursing Sertoli cells, the basal tight junctions of which form the blood testis barrier (BTB). The production of spermatozoa from spermatogonia through meiosis (i.e., spermatogenesis) lasts ~74 days in humans. Differentiating spermatogonia are generated every 16 days from stem spermatogonia, which also proliferate by mitosis to regenerate stem cells (508). Importantly, stem spermatogonia and primary spermatocytes up to pre-leptotene stage are not physically protected by the BTB from pathogens entering the interstitial tissue through blood vessels, since they are located in the outer wall of the tubule. The BTB opens in a tightly regulated manner to let germ cells progress towards the seminal lumen. Meiotic spermatocytes, postmeiotic spermatids, and spermatozoa located in the adluminal compartment are segregated by the BTB from interstitial components (e.g., viruses, immune cells, and antibodies). Sertoli cells phagocytose degenerated germ cells and residual bodies (remnants of spermatozoa cytoplasm) and play a key role for paracrine and endocrine regulations (e.g., inhibin synthesis), the latter under the control of pituitary folliculo-stimulating hormone (FSH) (145). Each seminiferous tubule is lined with myoid peritubular cells, the contractions of which help the release of spermatozoa in the tubule lumen. The epididymis is a single convoluted tubule (~5–7 m long in humans) structurally divided in three parts, head (caput), body (corpus), and tail (cauda), with distinct functions (e.g., absorption of testicular fluid and expulsion of spermatozoa during ejaculation). Only spermatozoa that have passed through the epididymis are mature enough to be capable of motility. Macrophages located in the lamina propria and in between the epithelial cells are the predominant epididymal immune cell types. Stromal T cells are mainly CD4+, whereas intraepithelial T lymphocytes are primarily CD8+. In humans, a small number of dendritic cells (DCs) are located mainly in the interstitial compartment of the epididymis head (153). The vas deferens is a narrow tube (~30 cm long in humans) with layers of smooth muscles that provide most of the propulsive force for ejaculation. Its distal part enlarges to form an ampulla, acting as a storage chamber. The vas deferens epithelium contains CD8+ T lymphocytes while CD4+ T lymphocytes and macrophages are found in the lamina propria (165). Each seminal vesicle consists of a single tube folded and coiled on itself. The seminal vesicles secrete a viscous alkaline fluid that provides nutrition for the sperm. The development, maintenance, and exocrine function of the seminal vesicles is highly dependent on testis-secreted androgens. The seminal vesicles contain stromal and intraepithelial macrophages, and to a lesser extent CD4+ (mainly in stroma) and CD8+ T cells (mainly within the epithelium). B lymphocytes are rare or absent (22, 141, 165). The prostate is a spherical tubuloalveolar exocrine gland regulated by male androgens and composed of complex branching glands in a fibromuscular lamina propria. The main function of the prostate is to secrete a slightly alkaline fluid that nourishes and protects the sperm and contains proteolytic enzymes (e.g., prostatic acid phosphatase, prostate-specific antigen) involved in liquefying the ejaculate. The prostate gland encompasses numerous immunocompetent cells including scattered T and B lymphocytes, macrophages, mast cells, and occasionally DCs in the stroma and epithelium. Foci of CD4+ T lymphocytes are frequent (22, 660, 683). The bulbourethral glands, also known as Cowper's glands, are small tubuloalveolar glands the secretions of which lubricate the ejaculate. A small population of scattered CD4+ T and B lymphocytes and macrophages is present in the connective tissue (464), and infiltrating CD8+ T cells are embedded in its epithelium.
IV. VIRUSES DETECTED IN THE MGT: ROUTES OF INFECTIONS AND TARGET ORGANS AND CELLS
The viruses that infect the MGT include both sexually acquired viruses responsible for localized genital infections and systemic viruses present in multiple organs and body fluids, which may or may not be sexually transmitted. For sexually acquired viruses leading to localized genital infections such as HSV and human papillomavirus (HPV), the penile mucosa is the main mode of entry into the MGT. Sexually transmitted viruses responsible for systemic infections such as HIV, HBV, and human T-lymphotropic virus (HTLV) can also infect men through the penile mucosa, which represents a major female-to-male mode of acquisition. However, their dissemination to internal MGT organs is believed to occur mainly through hematogenous spread, similar to other systemic viruses that are not transmitted through the penis, like ZIKV and EBOV.
Details of the routes to infection of the MGT used by a range of human viruses and a summary of the organs and cells they infect are as follows.
A. Sexually Transmitted Viruses that Infect the Penile Mucosa
The different penile mucosal epithelia (foreskin, shaft, glans, meatus, fossa navicularis, and penile urethra) represent a range of entryways for sexually transmitted viruses (FIGURE 5). Viruses can either infect the epithelial cells of the penile mucosa (e.g., HSV, HPV) or cross the penile epithelial barrier through a range of mechanisms that have been described in vitro (196) and partially in vivo (146, 206). These mechanisms, as for the ones occurring in the female tract, are detailed in FIGURE 11. The influence of hormones, microbiota, and genital fluids, described for viral infections of the female mucosa (81), are likely to modulate penile infection but there are no studies on this.

FIGURE 5.Anatomy of the human penis and immune cell distribution. The level of keratinization and thickness of the epithelia vary between the different penile mucosa (shaft, glans, inner and outer foreskin, urethral meatus, fossa navicularis) (19) and, along with mucus, constitute the first line of defense to viral invasions potentially transmitted during sexual intercourse. In circumcised men [where the foreskin is completely removed; ~30% of the male population worldwide (481)], the “wet” keratinized epithelium of the glans formerly covered by the foreskin becomes a dry epithelium and is believed to be more resistant to microabrasion during sexual intercourse, which in turn may limit viral exposure of target cells located within and below the epithelium layers. Multiple small urethral glands (called Littré glands) bud all along the epithelium of the penile urethra and secrete a mucous as a pre-ejaculatory emission, a lubricant thought to form an immunological barrier against invading pathogens due to the presence of secreted immunoglobulin A (IgA). Throughout the stratified squamous epithelia of the penis, there are numerous Langerhans cells (epidermal DCs) but few macrophages. T lymphocytes, in the majority of CD8+ cells, are found in the epithelia of the glans, foreskin, and meatus (188, 202, 444). In contrast to other parts of the penis, there are no DCs in the columnar urethra mucosa but many macrophages, as well as CD8+ T lymphocytes. The lamina propria of the foreskin encompasses predominantly macrophages, as well as memory CD8+ and a few CD4+ T lymphocytes including T helper 17 cells (19, 320, 555). The lamina propria of the penile urethra contains numerous T lymphocytes, essentially memory CD8+ T cells with a few CD4+ naive and memory T cells (most CD103+), natural killer (NK), memory B, and IgA and IgM producing plasma cells, but only a few macrophages (19, 618). In addition to the urethra, memory B cells and IgA and J chain + plasma cells are found in the glans and fossa navicularis mucosa, but are rare in the meatus (19, 505, 618). Compared with the urethra and fossa, the glans contains a higher number of NK cells that are more activated, as well as a higher number of effector CD8+ cells (618).
Of the 30-plus bacteria, parasites, and viruses responsible for sexually transmitted infections (STIs), the most common viruses that can be sexually acquired by men through penile infection are HPV, HSV, HBV, HIV, HTLV-1, and hepatitis D virus (HDV) (TABLE 2). Molluscum contagiosum virus (MCV), a viral skin infection often affecting the genital area, should also be included. Male acquisition of other hepatitis and herpes viruses upon sexual intercourse also occurs but is less frequent, and evidence of acquisition through the penis is either scarce [CMV, hepatitis C virus (HCV), Epstein-Barr virus (EBV), and Kaposi’s sarcoma associated herpesvirus (KSHV)] or nonexistent [hepatitis A virus (HAV) and hepatitis E virus (HEV)]. The transmission of emerging sexually transmissible EBOV and ZIKV through the penile mucosa is theoretically possible since they have been detected in cervico-vaginal secretions (475, 711), but never demonstrated.
Virus | Penile Acquisition | Penile Entryway | Detection in Penile Tissues | Risk Factor for Penile Infection | Effect of Circumcision | Nonsexual Transmission | Prevalence in Human Population | Effect on External Genitalia |
---|---|---|---|---|---|---|---|---|
HPV | Skin-skin contact with lesions; vaginal, oral, and anal sex (90) | Infection of basal cells in stratified epithelium (447, 655) | Shaft, glans, foreskin, and urethra (in vivo) (23, 158, 168, 630) | Urethritis, damaged epithelium (630) | Reduced acquisition, increased clearance (8, 686, 782) | Contact with lesions, vertical (598) | Europe: males 12.4−30.9% (280); USA: males 45.2%, females 39.9%; high-risk genital HPV males 25.1%, females 20.4% (450) | Genital warts, penile cancer for high-risk HPV types (79, 147) |
HSV-1/2 | Skin-skin contact with lesions; vaginal, oral, and anal sex (91) | Basal epithelial cells, capture by dendritic cells, cell-to-cell (1, 196) | Foreskin (607), glans and shaft (584), urethra (664) | HIV infection (>2.5-fold increase), genital ulcers, and penile epithelial trauma (455, 647) | Moderate reduction (28–30%) or no effect (647, 687) | Oral contact, contact with lesions, vertical (734) | 3.7 Billion/471 million (734) | Genital herpes, penile shaft, urethritis, prostatis, genital pain (604) |
HBV | Vaginal, oral, and anal (insertive over receptive) sex (314, 357) | Unknown mechanism | Not reported | Other STIs (82, 502) | Reduced risk of infection (82, 720) or no effect (MSM) (530) | Parenteral, vertical (314) | 257 Million (732) | |
HIV-1/2 | Vaginal, oral, and anal (receptive > insertive) sex (593) | Foreskin (inner >> outer), glans, coronal sulcus, fossa navicularis, and urethra (19) | Glans, foreskin, urethra (200, 298, 776) | Other STIs (701, 735), penile microbiome and anaerobic dysbiosis (396, 605), penile inflammation (19, 556) | Circumcision reduces acquisition by 56–61% (174) | Parenteral, vertical (735) | 36.9 Million (735); 51% are female (701) | |
HTLV-1 | Vaginal and anal sex (526) | CD4+ T cells, virions capture by dendritic cells (311) | Not reported | Syphilis (235, 491), HSV-2 (246), potentially other STIs (526) | Unknown | Parenteral, vertical (311) | 5–10 Million (222) | Erectile dysfunction (520, 521) |
HTLV-2 | Vaginal and anal sex (585) | CD4+ T cells, virions capture by dendritic cells (311, 585) | Not reported | Potentially other STIs (235) | Unknown | Parenteral (612), vertical (58) | Not estimated in the global population | Erectile dysfunction (552) |
MCV | Skin-skin contact with lesions; vaginal, oral, and anal sex (131) | Lower and outer layers of the epidermis (131, 568) | Glans, shaft, foreskin (568) | HIV, immunosuppression (92, 266) | Not reported | Fomites, contact with lesions (131) | Worldwide: 2–8%; HIV-infected: 5–18% (309) | Skin-colored papules (131) |
HDV | Vaginal and anal sex (17) | Unknown mechanism | Not reported | HBV infection (733) | Unknown | Parenteral, vertical (733) | 15–20 Million (733) |
1. Viruses responsible for localized genital infection
a) human papillomaviruses.
Over 40 types of HPV are sexually transmitted and infect the genitals or anus. HPV is the most common infection of the reproductive tract, with over 80% of sexually active adults infected by one HPV at least once in their lifetime (105). HPV infections are usually cleared by the immune system within a few months of acquisition. However, a small proportion of infections with certain types of HPV can persist and progress to cancer, including cervical cancer (the most prevalent) and penile cancer. HPVs are classified as either high risk (at least 14 types, e.g., HPV 16, 18, etc.) or low risk (e.g., HPV 6, 11, and others) based on their presence or absence in cervical cancer. Interestingly, two oncogenic HPV lineages were introduced in our modern ancestors through sexual intercourse with infected Neanderthals/Denisovan populations (102, 546).
HPV infects the basal cells of stratified epithelium, the only tissue in which it replicates, through breaks in the epithelial barrier integrity, such as micro-abrasions caused by sexual intercourse (655). Virions enter the cells by endocytosis after binding to putative receptors (α integrins, laminins, and annexin A2) on the basement membrane (567). Antibodies play a key role in neutralizing the virus while it resides on the basement membrane. HPV does not kill the infected cell and remains as a chronic asymptomatic infection, with episomal copies of its genome in the cell nucleus. Viral particles are released with desquamating cells. Although the virus is transmissible as soon as it infects the basal epithelial cells, it takes months or years before squamous intraepithelial lesions develop and can be clinically detected. HPV has been found on the scrotum and external parts of the penis (glans, shaft, foreskin), and with a lower prevalence in the urethra, ductus deferens, epididymis, and testis (158). In a cohort of Japanese men with urethritis, HPV prevalence was similar to that in the glans (20% as opposed to 31%) (630), possibly due to increased acquisition through the already damaged urethral epithelium, or to enhanced detection because of the increased exfoliation of infected cells. Male circumcision decreases the prevalence of HPV in men, including high risk (782), and has been associated with reduced acquisition of the virus (when scrotum and penis shaft sampling were excluded) (9, 253) as well as with increased viral clearance (253, 286, 686). Overall, these data suggest that the foreskin constitutes a favorable environment for HPV infection.
b) herpes simplex viruses.
HSV-1 and -2 induce lifelong infection in humans. HSV-1 is mostly transmitted by oral-to-oral contact, causing oral herpes, but can also be transmitted by sex and cause genital herpes. In the last decade, an increasing proportion of genital herpes has been attributed to HSV-1 in developed countries, probably arising from improvements in hygienic standards. This is evidenced by the acquisition of the HSV-1 infection later in life, during the initiation of sexual activity, rather than in childhood (753). Moreover, the increasing acceptability of oral-genital sex favors the acquisition of genital HSV-1 (579). Genital HSV-1 is most prevalent in America, Europe, and the Western Pacific. Globally, 140 million adults (~25% of the people infested with HSV-1) were estimated to have genital HSV-1 infection in 2012, but prevalence varied substantially by region (744). HSV-2, which infects an estimated 471 million people under the age of 50 worldwide (64% women), is the main agent of genital herpes and is transmitted almost exclusively via vaginal and anal sex. Similar to other pathogens such as HTLV-1 and HIV, HSV-2 is more efficiently transmitted from men to women than from women to men (744). Risk factors for HSV-2 acquisition in men include HIV infection (2.5- and >3-fold increase in 2 trials) (455, 647), genital ulcer disease, and penile epithelial trauma (455). How HIV infection facilitates HSV-2 acquisition is unclear, but might involve modification of the genital immune protection against HSV-2 caused by HIV, or shared risky behavior.
HSV (either 1 or 2) first replicates in keratinocytes in mucosal surfaces and subsequently infects nerve cells via nearby nerve endings and retrograde axonal transport, where it establishes latent infection. Upon reactivation, HSV is transported back to the epithelium, creating lesions due to virus-induced cell death (676). The cell-surface, heparan sulfate proteoglycans, widely expressed by epithelial cells, are involved in the attachment of HSV to the target cells, but viral entry only occurs upon recognition of specific cellular co-receptors (654). The co-receptor(s) used by HSV to infect the penis mucosa are unknown. In the foreskin, HSV-2 did not colocalize with nectin-1 (607), one of the most frequently studied HSV receptors. In two trials conducted in Africa involving ~3,000 men each, medical male circumcision resulted in a moderate reduction (28–30%) (647, 687) in male HSV-2 acquisition, while in a third African trial involving ~2,000 men, no reduction was observed, even at a 6-yr follow-up (455, 456). The results of these trials suggest that the foreskin does not represent the main entryway for HSV. In vitro studies on reconstructed foreskin epithelial layers and monolayers of polarized keratinocytes showed that HSV-1 preferentially infects the basal layer of the epithelium (607, 715), and requires accessibility of cellular basolateral membranes to infect, whereas cells that only expose the apical surface are resistant to viral entry (607). This explains why damage to the genital epithelium integrity caused by ulcers or trauma (e.g., micro-abrasions occurring during sexual intercourse) favor HSV acquisition, whereas an intact epithelium offers relative protection.
The occurrence of urethritis in HSV-1-infected men suggests that the urethra might represent a site of entry for HSV-1 (70). In addition to epithelial cells, HSV-1 and HSV-2 infect human dendritic cells (DCs) (64). Opsonization by blood-derived DCs of HSV-2 particles coated with complement molecules that are naturally present in genital secretions enhanced the productive infection (124). Infection of DCs might represent an alternative mechanism of HSV penetration into the penile mucosa, and a way for the virus to escape immune recognition. The activation of virus-specific cytotoxic T lymphocytes is central to the immune control of HSV (217, 392, 631, 778) and requires presentation of viral antigens by DCs in secondary lymphoid organs. Infection of DCs by HSV in mucosal tissues might impair their ability to migrate to lymphoid organs and present viral antigens to T cells, as shown for HSV-1 in the skin (260). In addition, the cell-to-cell spread of HSV through viral synapses (260) allows the virus to hide from neutralizing antibodies (603) and may contribute to penile infection. Genital ulcerative disease caused by HSV is associated with increased risk of HIV acquisition (606, 685).
2. Viruses responsible for systemic infections
a) human hepatitis b virus.
Sexual intercourse is the most common way that HBV is disseminated in areas of low to intermediate prevalence of infection (314). Sexual transmission is most common in MSM (men having sex with men), but HBV is also readily transmitted through heterosexual intercourse. One study identified insertive rather than receptive anal intercourse as the major risk factor for HBV seroconversion, suggesting that penile acquisition is a particularly efficient mode of transmission (357). Male circumcision decreases the risk of HBV acquisition, while HIV-1 seropositive status is a risk factor (530, 720). The main target cells for HBV are the hepatocytes, but the virus also infects lymphoid cells (663), detected in early studies; it was detected in the testis and other organs (438). Despite its transmission being much more efficient than that of HIV (357), the acquisition mechanisms for HBV through the genital mucosa have not been investigated, mainly due to the lack of robust cell culture systems for this virus (710). Similarly, the transmission mechanisms for HDV, which requires HBV for replication and is acquired through the same transmission routes, are unknown.
b) human immunodeficiency virus.
Most cases of HIV infection worldwide (~2 million new infections each year) are the result of sexual transmission, with greater risk for male-to-male than heterosexual transmission (593). Indeed, receptive and insertive vaginal intercourse risk of HIV acquisition is estimated at ~8 and 4 infections per 10,000 exposures, respectively. Risk of penile acquisition of HIV by insertive anal sex is ~11 infections per 10,000 exposures. The highest risk occurs in people practicing receptive anal intercourse (138 infections per 10,000 exposures) (534). The main cellular targets for HIV are the cells of the immune system, namely, CD4+ T lymphocytes, cells of the macrophage lineage, and some population of DCs. The receptor for viral entry is the CD4 molecule, and the C-C and C-X-C motif chemokine receptors CCR5 and CXCR4 function as main co-receptors. The foreskin represents an important entry route for HIV, as demonstrated in clinical trials by lower incidence/prevalence following circumcision (a reduction of 56–61%) and epidemiological studies (19, 34, 36, 174, 252). As a result of these reports, the World Health Organization (WHO) recommended male medical circumcision as an important element in HIV prevention programs (731). Although these studies point at the foreskin as a highly susceptible site of infection, they also indicate that HIV can access the MGT, and in turn the systemic circulation, through the infection of other penile mucosal epithelia. Studies on HIV infection of various penile sites has demonstrated that explants of inner foreskin, glans meatus, and urethra were all susceptible to HIV-1 infection (188, 202, 535). When a polarized model of foreskin explant was developed to more closely mimic the physiological exposure of the apical side of the epithelium to incoming virus, efficient HIV-1 transmission occurred in the inner foreskin but not the outer foreskin following exposure to infected cells, whereas cell-free virus did not infect. Cell-mediated infection required the formation of viral synapses between mononuclear infected cells and apical foreskin epithelial cells, internalization of HIV-1 by epidermal Langerhans cells, and transfer of the virus to T cells by Langerhans cells (202). Exposure of the inner foreskin to HIV-1-infected cells resulted in RANTES-mediated recruitment of T cells into the epidermis [RANTES being Regulated on activation, normal T cell expressed and secreted or chemokine (C-C motif) ligand 5 (CCL5)] and CCL20-mediated trafficking of Langerhans cells (776). In ex vivo exposed glans and foreskin tissue, more viral particles entered the glans penis as compared with foreskin tissue, and to greater depths (146). However, after 24 h of culture, HIV particles were primarily visualized in the inner side of the foreskin, similar to observations in vivo in Rhesus macaque upon penile inoculation of SIV (146). Polarized human urethral explants demonstrated that CCR5+/CD4+ macrophages were the initial target of HIV-1 upon exposure of penile tissue to infected cells, while exposure to cell free virus, again, did not result in infection. The fossa navicularis and glans appeared relatively resistant to HIV-1 (201). A study in Rhesus monkeys showed that the T cells, macrophages, and DCs of the glans, foreskin, and coronal sulcus are the primary targets of SIV, after penile inoculation of the virus. SIV next reached the genital lymph nodes and disseminated to the bloodstream and lymphoid system within 1 wk. After 14 days, SIV had contaminated all the tested tissues (411). A model of repeated SHIV (a chimeric simian-human immunodeficiency virus containing HIV envelope gene) penile exposure of Rhesus macaques further suggested that the urethra is more susceptible to HIV acquisition than is the inner foreskin or glans tissue (206). As with other STIs, the mucosal acquisition of HIV/SIV is influenced by the genital microbiota (174), immune response (505), and host genetic factors (399, 651, 759). In addition, the risk of HIV acquisition is increased by other STIs such as HSV-2 (~3-fold increase), most likely due to disrupted genital mucosae as a result of ulcerative STIs, and the inflammation-driven recruitment and activation of HIV target cells residing in the penile tissues (174).
c) human t-lymphotropic virus type 1/2.
Sexual acquisition is the second most common mode of transmission of HTLV-1 in endemic areas. Like HIV, HTLV-1 is transmitted more efficiently from males to females than from females to males (526). The mechanism of HTLV penile transmission is unknown. Considering the exclusive tropism of HTLV-1 for T lymphocytes and the requirement of cell-to-cell contact for infection (198), it may involve direct contact between infected T cells in female genital secretions and T cells present in the penile mucosa and/or internalization of cell-free particles by DCs and subsequent transfer to lymphocytes by cell-to-cell contact (327). Female-to-male transmission is higher in men with syphilis or a history of penile sores or ulcers (491), consistent with the concept that disrupting the genital epithelium increases viral transmission. A closely related virus is HTLV-2, distributed mainly among native American population (589). The mode of transmission and the target cells (T lymphocytes) of HTLV-2 resemble that for HTLV-1, but is generally asymptomatic.
B. Viral Infections of Internal Male Genital Organs
1. Hematogenous and nonhematogenous routes of infection
For systemic viruses that generate high viremia, e.g., during the acute stage of infection of HIV, EBOV, and ZIKV, hematogenous spread probably represents the main entry portal into the MGT. Viruses may then enter the lamina propria and testis interstitium through either passive diffusion of viral particles across the vessels, endothelial cell infection, or active passage of infected cells. The biphasic infection process observed for ZIKV within the mouse testis and epididymis, i.e., infection of cells located close to blood and lymphatic vessels before seminiferous tubules and epididymal epithelial cell spreading, suggest partial protection of these compartments by the myoid cells and extracellular matrix layers (441, 696). Of note is that the barrier represented by the myoid peritubular cell layers in rodents’ testis does not exist in humans.
Virus transfer between the testis and epididymis has been recently suggested by phylogenetic analysis of viral strains in SIV-infected macaques (299) and by micro-RNA targeted viral clones for ZIKV in immune-deficient mice (696). Migration to the epididymis of infected cells such as germ cells, rete testis macrophages, and/or connected vasculature could be involved. In addition, downstream MGT ducts and glands might be infected by the viral particles or infected cells carried by the excurrent flow during ejaculation; phylogenetic analyses in SIV-infected macaques suggested that urethral viruses originated partly from viruses present in semen. Reinfection of the urethra by seminal viruses might be facilitated by the stagnation of residual semen in the penile urethra post-ejaculation (299).
2. Targeted cells and tissues
Viruses belonging to a broad range of families have been found in the organs of mammalian MGTs, including humans (FIGURE 6), by means of polymerase chain reaction, in situ hybridization, electron microscopy, immuno-detection of viral antigens, and viral isolation (135, 138). TABLES 3 and 4 present an updated list of the viruses detected in the male genital organs and cells. Depending on virus tropism, various cellular populations of the male organs are targeted, including epithelial cells (e.g., HPV), endothelial cells (e.g., EBOV), fibroblasts (e.g., CMV), immune cells (e.g., HIV, CMV), Leydig cells [e.g., mumps virus (MuV)], Sertoli cells [e.g., Marburg virus (MARV) and ZIKV], and germ cells [e.g., ZIKV, human herpesvirus 6 (HHV-6)]. Although the tight junctions of Sertoli cell protect the testicular meiotic and post-meiotic germ cells from pathogens in the interstitium, several viruses have evolved mechanisms to hijack tight junction proteins by either using them as receptors or co-receptors for cell entry, or by opening them to facilitate dissemination (501, 689). In addition, as suspected for ZIKV, viruses can bypass the testicular Sertoli cell barrier through a variety of mechanisms, including 1) direct infection of Sertoli cells (441), 2) viral transcytosis (i.e., passage through the cell cytoplasm without infection) from the Sertoli cell basal to apical pole (634), and 3) disruption of the epithelial barrier induced by inflammatory mediators released by infected cells in the vicinity (634, 665).

FIGURE 6.Viruses detected in human male genital tract in vivo. The figure recapitulates the viruses detected in human biopsies or secretions of the internal genitalia. HPV, human papillomavirus; HSV, herpes simplex virus; EBV, Epstein-Barr virus; CMV, cytomegalovirus; HHV, human herpesvirus; KSHV, Kaposi sarcoma-associated herpesvirus; HIV, human immunodeficiency virus; HBV, hepatitis B virus; MuV, mumps virus; AAV, adeno-associated virus; ZIKV, Zika virus; WNV, West-Nile virus; EBOV, Ebola virus; MCPyV, Merkel cell polyomavirus; JCPyV, JC polyomavirus; BKPyV, BK polyomavirus; SV40, simian virus 40.
Human | Animal Models | ||||
---|---|---|---|---|---|
Virus | In vivo | In vitro | NHPs | Rodents | Reference Nos. |
HPV | Int, SC | 435 | |||
HSV-1 | Int, SC, GC | 423 | |||
HSV-2 | Specific tropism not defined | GC | 143, 256 | ||
EBV | Specific tropism not defined | 106 | |||
CMV | LC, GC, F | 500 | |||
HHV-6 | Endogenous | 480 | |||
AdV | Specific tropism not defined | Specific tropism not defined | 127 | ||
Parvovirus B19 | Specific tropism not defined | 149, 170, 251 | |||
AAVs | Specific tropism not defined | LC, PT, SC, GC (Sptg) | 171, 454, 726 | ||
HBV | F, End | 378, 438 | |||
HIV-1 (-2) | Mϕ, T cells, GC | Mϕ, T cells | SIV: Mϕ, T cells, GC | 742 | |
HERVs | Endogenous | 125 | |||
MuV | Specific tropism not defined | LC | LC, SC, GC, Mϕ | 59, 244, 748 | |
ZIKV | GC* | LC, SC, GC, Mϕ, PT | Specific tropism not defined | LC, SC, GC, Mϕ, PT | 441, 657 |
WNV | Within ST, SC | SC | 28, 634, 645 | ||
JEV | Specific tropism not defined | 682 | |||
EBOV | End, within ST | End, GC | Int | 113, 430, 541, 766 | |
MARV | SC, GC, Mϕ, F, End | 111, 117 | |||
CoxV B5 | Specific tropism not defined | 122 | |||
Poliovirus | Specific tropism not defined | Specific tropism not defined (in vitro) | 670 | ||
HEV | GC (Sptg) | Within ST | 303, 652 | ||
SARS | LC, Ep | 770 | |||
LCMV | Int, GC (Sptg) | 179 |
Virus | Epididymis | Vas Deferens | Seminal Vesicles | Prostate |
---|---|---|---|---|
HPV | Viral DNA (669) | Viral DNA (573, 669) | Viral DNA in seminal plasma of vasectomized men (573) | Viral DNA (101, 448, 449, 573, 642, 729, 749, 764) |
HSV-2 | Viral isolation in guinea pig (661) | Viral isolation in guinea pig (661) | Viral isolation from biopsies (143) | Viral DNA (95, 749) |
EBV | Viral DNA (101, 259, 729) | |||
CMV | Detection in epithelium (356) | Detection in epithelium (13, 356) | Viral DNA (101, 600, 749, 761) | |
KSHV | Viral DNA (285, 473) | |||
JCPyV | Viral DNA (101, 642, 764) | |||
BKPyV | Viral DNA (101, 764) | |||
MCPyV | Viral DNA (642) | |||
SV40 | Viral DNA (101) | |||
HIV | Viral RNA and antigens in leukocytes (511, 557) | NHP models: viral RNA and antigens in leukocytes (299, 440) | Viral RNA and antigens in leukocytes (141) | Viral RNA and antigens in leukocytes (114, 511, 557, 644) |
WNV | Viral antigens detected in biopsies (28) | |||
CoxV A6 | Viral RNA in epididymal fluids (719) | |||
ZIKV | Mouse models: in mature sperm, epithelial cells, and stroma (96, 110, 151, 157, 247, 257, 345, 445, 702, 703) | Mouse model (345) | Mouse model: in epithelial cells and stroma (110, 157, 445); NHP model (287, 524) | Mouse model: in epithelial cells and stroma (96, 110); NHP model (287, 524) |
EBOV | NHP model: in stromal cells, endothelial cells, and macrophages (541, 766) | NHP model: in stromal and endothelial cells (541) | NHP model: in stromal and endothelial cells (541) | |
MARV | NHP model (117) | NHP model (117) | ||
HEV | NHP model (303) |
V. MGT SENTINELS, WEAPONS, AND ACHILLES HEEL IN RELATION TO VIRAL ATTACKS
There follows a review of the little data that are available on humans and the bulk of studies on rodents relating to the immune specificities of male genital organs and their ability to respond to viral attacks. FIGURES 4, 5, and 7 show the localization of specialized immune cell types in MGT organs, and TABLE 5 shows the expression of pathogen sensors throughout the MGT.

FIGURE 7.Immune privilege in the testis. Interstitial resident macrophages are the main immune cell population in the testis. A few T lymphocytes (mostly central memory CD8+, regulatory, and natural killer T cells), myeloid dendritic cells (mDC), mast cells, natural killer (NK) cells, and eosinophils are also naturally present in the interstitial tissue, whereas B lymphocytes and neutrophils are absent (322, 550, 771). Interstitial testicular macrophages are closely associated physically and chemically with Leydig cells and play an important role in their development and regeneration and in testosterone production (55, 485). In the rat testis, a subpopulation of testicular macrophages of embryonic origin closely associated with seminiferous tubules (so-called peritubular macrophages) was characterized (134, 484). Besides the Sertoli cell tight junctions forming the blood testis barrier (BTB), the immune privilege of the testis is maintained by immunosuppressive factors secreted by Sertoli cells (SIF), interstitial macrophages, mDCs, and T lymphocytes [e.g., transforming growth factor-β (TGF-β), interleukin (IL)-10, activin A, indoleamine-pyrrole 2,3-dioxygenase (IDO), galectin-1], which suppress T cell activation and favor Treg expansion. Growth arrest-specific gene 6 (Gas6)/protein S (ProS) signaling through TAM receptors inhibits the Toll-like receptor (TLR)-initiated innate immune response in Sertoli and Leydig cells. Testosterone secretion by Leydig cells regulates cytokine expression in Sertoli and peritubular cells, consolidates Sertoli cell tight junctions, and controls the proportion of testicular leukocytes. Germ cell autoantigens arising from apoptotic germ cells/residual bodies are either phagocytosed by Sertoli cells under TAM receptor regulation, or presented by peritubular macrophages to Treg, in turn preventing inflammatory responses. Treg, regulatory T cells; SC, Sertoli cells; LC, Leydig cells; P, peritubular cells; iMϕ, interstitial macrophage; pMϕ, peritubular macrophage; CM T lymphocyte, central memory T lymphocyte; TAM-Rc, TAM receptors.
Testis | Epididymis | Prostate | ||||||||
---|---|---|---|---|---|---|---|---|---|---|
PRR | PRR Ligand and Virus Sensed | Rodent | Human | Rodent | Human | Rodent | Human | Urethra (Human) | Foreskin (Human) | |
TLRs | TLR2 | Surface viral gp, MuV, CMV, HSV | Sertoli cells (572, 747) | Sertoli cells, peritubular cells, and a subset of interstitial cells (376, 442) | Strongly expressed by epithelial cells except clear cells (528) | ND | Epithelial cells (215, 414) | mRNA but (−) for protein expression (510) | Lymphocytes (558) | Keratinocytes (777) |
TLR3 | Cytosolic dsRNA, KSHV | Sertoli cells (656, 747), Leydig cells (620), macrophages (56), spermatogonia and spermatocytes (302, 724) | Sertoli cells (376, 442) | Epithelial cells except clear cells (528, 781) | Epithelial cells (cauda > caput cells) (77) | ND | Epithelial cells in a subset of prostate samples (559) | Lymphocytes (558) | Keratinocytes (777) | |
TLR4 | Surface viral gp, HIV-1 gp120 | Sertoli cells (572, 747), Leydig cells (620), macrophages (low level) (56) | ND | Strongly expressed by epithelial cells except clear cells (528) | ND | Epithelial cells and stromal cells (215, 414, 565) | Epithelial and stromal cells (214) | Lymphocytes (558) | Keratinocytes (777) | |
TLR7 | Cytosolic ssRNA, SeV, VSV | ND | ND | Strongly expressed by epithelial cells except clear cells (528) | mRNA poorly expressed (77) | ND | Intraepithelial macrophages (559) | Intraepithelial macrophages (558) | ND | |
TLR8 | Cytosolic ssRNA | ND | ND | Strongly expressed by epithelial cells except clear cells (528) | mRNA poorly expressed (77) | ND | Epithelial cells in a subset of prostate samples and intraepithelial macrophages (559) | (−) | ND | |
TLR9 | Cytosolic DNA, HSV-1/2, KSHV | ND | ND | Strongly expressed by epithelial cells except clear cells (528) | mRNA poorly expressed (77) | ND | Intraepithelial macrophages (559), epithelial and stromal cells in BHP (312, 705) | Epithelial cells and intraepithelial macrophages (558) | Keratinocytes (777) | |
RLRs | RIG-1 | Cytosolic dsRNA, MuV, ZIKV, SeV | Leydig cells (779) | ND | Epithelial cells (781) | Epithelial cells (cauda > caput cells) (77) | ND | ND | ND | ND |
MDA-5 | Cytosolic dsRNA, DENV, WNV | Leydig cells (779), spermatids (779) | ND | ND | Epithelial cells (cauda > caput cells) (77) | ND | ND | ND | ND | |
LGP2 | Cytosolic dsRNA | ND | ND | ND | ND | ND | ND | |||
NLRs | NLRP3 | Cytosolic ssRNA | ND | ND | ND | ND | Rat prostatic cells (98) | ND | ND | ND |
NOD2 | Cytosolic dsRNA | ND | ND | ND | ND | Mouse epithelial cells (330) | ND | ND | ND | |
ALRs | cGAS | Cytosolic DNA, HIV, KSHV | ND | Epithelial cells, following HSV-60 stimulation (781) | ND | ND | ND | ND | ND | |
ZBP1/DAI | Cytosolic DNA, CMV | ND | DAI: epithelial cells (781) | mRNA ZBP1 in epithelial cells (cauda > caput) (77) | ND | ND | ND | ND | ||
p204/IFI16 | Cytosolic DNA, KSHV | p204 in Leydig cells (780) | ND | p204 in epithelial cells following HSV-60 stimulation (781) | mRNA IFI16 in epithelial cells (cauda > caput) (77) | ND | IFI16 in epithelial cells (750) | ND | ND | |
CLRs | CD206 | ND | ND | ND | ND | ND | Macrophages | ND | ND |
A. The Testis
1. An immune privileged organ
The testis possesses specific immunity characteristics to prevent immune rejection of the developing germ cells and spermatozoa that appear during puberty, following the establishment of systemic self-tolerance, and which carry high surface antigen load. The initial discovery of the testis as an immune privileged site came from the observation that allografts in rat testis enabled prolonged survival (342). However, allografts in the testis from larger animals, including non-human primates (NHPs), showed a wide variation in immune protection duration (342), suggesting species-specific characteristics. Rodent studies showed that immune privilege in the testis is dependent on 1) physical segregation by the blood-testis barrier (BTB) of the postmeiotic spermatids and meiotic spermatocytes localized in the adluminal compartment of the seminiferous tubules and 2) an active local immunosuppressive and tolerogenic environment (27, 100, 771). Indeed, the BTB is insufficient to sequester all germ cell autoantigens, since some of the meiotic germ cells (i.e., primary spermatocytes up to the preleptotene stage) are located in the basal compartment of the seminiferous epithelium, in reach of immune components. Below is a summary of the cellular and molecular basis for the testis immune privilege, mostly established in rodents, along with data for human and NHP (FIGURE 7).
a) role of resident immune cells in the immune privilege of the testis.
In both rodents and humans, macrophages represent the largest population of immune cells in the testis and play a key role in maintaining testis immune privilege (457, 485).
Rodent and primate testis macrophages display an M2-like phenotype. In this regard, they express the scavenger receptor CD163 (550, 586, 739), secrete high levels of immune-suppressive interleukin-10 (IL-10) and transforming growth factor-β (TGF-β), and low levels of pro-inflammatory tumor necrosis factor-α (TNF-α) (586, 739). In response to infection or inflammatory stimuli, rodent testis macrophages exhibit relatively low inflammatory responses and high immunosuppressive properties compared with macrophages located in others tissues (reviewed in Ref. 457), which confer upon them a unique immune-tolerant role against meiotic germ cells antigens that egress from seminiferous tubules.
Blocked nuclear factor κB (NF-κB) signaling pathway and basal corticosterone production in rat testicular macrophages are associated with reduced pro-inflammatory capacities (457). In the mouse, interstitial embryo-derived macrophages have more potent immune suppressive properties than peritubular macrophages, which are derived from the bone marrow after birth (484). Rodent and human testis interstitium contain a small population of DCs, mostly of myeloid phenotype (538, 550), which in rats are unable to induce T cell proliferation under physiological conditions. They release anti-inflammatory cytokines [e.g., IL-10, indoleamine 2,3-dioxygenase (IDO)] and induce T regulatory cells (Tregs) instead of autoreactive T cells following antigenic presentation (263, 723). Low numbers of T lymphocytes, essentially central memory, are present in human and animal testis (CD8+ T cells mainly, with a few CD4+ T cells, Tregs, and natural killer T cells) (586, 688). Their response to foreign antigens tends to be inhibited (586), probably because of local immunosuppressive factors (e.g., IL-10, TGF-β, activin A, lyso-glycerophosphocholine) (193, 282). Immunosuppressive purinergic signaling was reported in human testicular Tregs and memory CD8+ T cells (322). Egress of spermatid fragment antigens during spermiation contributes to mouse testis immune tolerance (698), possibly through the presentation of autoantigens to Tregs by major histocompatibility complex II (MHC II) positive peritubular macrophages (485).
b) role of non-immune testicular cells.
The testosterone produced by Leydig cells balances pro- and anti-inflammatory cytokine expression in Sertoli cells and peritubular cells (184), regulates testicular macrophage and lymphocyte numbers and pro-inflammatory properties, and stimulates Tregs expansion (184, 186). Androgen signaling in Sertoli cells is also essential for BTB integrity (459, 460). Leydig cells also produce growth arrest-specific gene 6 (Gas6)/ProS, which inhibits innate immune response in DCs, macrophages, and Sertoli cells (620). Rodent and human Sertoli cells express several immunoregulatory factors (TGF-β, IDO, galectin-1, activin A) that induce Treg and tolerogenic DCs (205, 343). The co-transplantation of Sertoli cells with tissue in mice and NHPs prolonged graft survival (404). Phagocytosis of apoptotic germ cells and residual bodies by Sertoli cells under TAM receptor regulation avoids inflammatory responses to damaged germ cell auto-antigens (751). Fas ligand (FasL), originally described on Sertoli cells, was subsequently found on germ cells and might induce T cell apoptosis (47, 128). The programmed death-1 (PD-1)/PD-1 ligand (PDL-1), expressed by spermatocytes and spermatids system, which promotes the development and function of Tregs and suppresses effector T cell-mediated immune response, is also suspected of contributing to testis immune suppression (103).
2. Testicular immune responses to invading virus
The sensing of a pathogen by pattern recognition receptors (PRRs) (TABLE 5) expressed by both immune and nonimmune cells is key in initiating antiviral responses. Upon activation, these sentinels send intracellular signals that lead to the expression of innate immune mediators such as antiviral interferons (IFNs) and pro-inflammatory cytokines. Additionally, PRR-triggered cell signaling induces several nontranscriptional processes such as phagocytosis, autophagy, cell death, and inflammasome/cytokine processing. In short, the innate immune system can counteract the infection through 1) the elimination of virus-infected cells by macrophages, neutrophils, and natural killer cells or by apoptosis; 2) the production of proteins with antiviral activity [Upon binding to their receptor, IFNs induce the synthesis of hundreds of interferon-stimulated genes (ISGs) that inhibit different steps of viral replication (611)]; and 3) the mobilization of adaptive immune response through pro-inflammatory cytokine/chemokine production.
a) rodent models.
Rodent somatic testicular cells appear well-equipped to defend against viral attacks. Rat Leydig cells and Sertoli cells produced large amounts of type I IFN, ISGs, and pro-inflammatory cytokines when exposed in vitro to Sendai virus, an RNA virus from the same viral family as MuV (136, 137, 139, 243, 458). MuV induced antiviral responses in mouse Sertoli and Leydig cells through Toll-like receptor 2 (TLR2) and retinoic-acid inducible gene I (RIG-I) signaling (746). Sertoli cells also expressed functional TLR3 and TLR4 (572, 656, 747). However, TLR-initiated innate immune response in Sertoli and Leydig cells were negatively regulated by Gas6/ProS-TAM signaling, which may prevent a sustained inflammatory response (620, 667). PolyIC, an agonist of the RNA virus sensors TLR3, RIG-1 and melanoma differentiation-associated protein 5 (MDA5), disturbed testosterone secretion by Leydig cells, whereas HSV-60, an agonist of the cytosolic DNA sensor p204, had no effect, despite p204 expression (620, 779, 780). Exogenous type I IFN induced ISGs and inhibited MuV replication in Leydig cells, macrophages, and Sertoli cells (748), demonstrating effective IFN signaling cascade. MuV-exposed Sertoli cells secreted high levels of pro-inflammatory cytokines, among which chemokine (C-X-C motif) ligand 10 (CXCL10) triggered germ cell apoptosis (325), suggesting that Sertoli cells might play a key role in MuV-induced inflammation and germ cell damage in the testis (746).
Rat testicular macrophages expressed lower levels of TLR4 receptors than peritoneal macrophages (56) as well as low basal expression of TLR-signaling pathway genes (e.g., CD14, TRIF, TRAF6, IRAK1, TAK1, etc.), while the negative regulation of TLR-signaling pathways such as IkBα, SARM, and RP105 were highly expressed (56). Nevertheless, testicular macrophages were able to display IFN, ISGs, and pro-inflammatory cytokine production following Sendai virus exposure (139, 140, 243).
In contrast, Sendai virus (SeV) (136), MuV (746), and ZIKV (581) infection did not trigger an IFN-based antiviral response in differentiating rat and mouse germ cells. Nevertheless, poly IC induced the production of pro-inflammatory cytokines and type I IFN through TLR3 and MDA5 signaling in mouse germ cells, but to a much lower level than in Leydig cells (302, 724). Exogenous IFN stimulation failed to induce a strong upregulation of ISGs in differentiating testicular germ cells, which is in line with the lack of functional type I IFN receptor expression in meiotic and post-meiotic mouse germ cells (136, 581, 602). However, male germ cells are equipped with autophagic machinery (621), and autophagy contributed to blocking MuV replication in mouse germ cells and testicular macrophages (748). Additionally, the stem germ cell spermatogonia differed from their daughter cells in that they expressed both subunits of the type I IFN receptor and produced low levels of ISGs upon viral stimulation (136, 602). Interestingly, overexpression of type I IFN in mouse testis triggered germ cell apoptosis and induced sterility. This effect might be mediated by IFN-induced alterations in Sertoli cells or spermatogonia (602).
b) humans and nhps.
In contrast to their rodent counterparts, human Leydig cells infected by the MuV or stimulated by poly IC in vitro did not upregulate type I or II IFNs and only slightly upregulated ISG expression, whereas pro-inflammatory CXCL10 was highly expressed (244, 692). Acute infection of a human Sertoli cell line by ZIKV induced an upregulation of genes implicated in innate immune response and IFN signaling, but this response was downregulated after prolonged infection (634, 665). In human testis tissue exposed to ZIKV ex vivo, the virus infected a broad range of somatic and germ cell types and triggered the transcription of a number of ISGs, but surprisingly did not stimulate type I, II, or III IFN transcripts or proteins. Additionally, the pro-inflammatory response was restricted to CXCL10 upregulation (441). This muted innate response is unlike that observed in mouse testis infected with ZIKV in vivo (441, 657). It might explain the absence of orchitis in ZIKV-infected men and contribute to viral persistence in the human testis (see sect. VIC). The apparent lack of IFN upregulation in the infected human testis requests further investigation. As for human peritubular cells, they produced IL-6 and monocyte chemoattractant protein 1 (MCP-1) upon activation of TLR2 (442). In MARV-infected NHP, the persistent infection of Sertoli cells was associated with BTB disruption, germ cell depletion, and leukocyte infiltration with numerous Tregs, the latter postulated to prevent virus clearance (111). Similarly, testis inflammation and lesions, along with testis pain and orchitis, were observed in MARV-infected patients (609). In contrast, EBOV, another filovirus with testicular tropism, did not induce any apparent damage in human and NHP testis (609). Differences in the infected cell types and in the PRRs activated by viruses in the testis most probably have a strong influence on the testicular immune responses and infection outcome.
In SIV-infected macaques, testis infection did not lead to leukocyte infiltration or elevated cytokine levels, unlike prostate infection (625, 693, 740). Although an increase in the relative proportion of effector memory CD8+ T cells was detected in the testis of SIV-infected pigtailed macaques, these cells had suppressed cytokine responses to mitogen activation (740). In the testis of HIV-infected individuals, elevated CD4+ and CD8+ T cell immune activation was postulated to favor HIV replication (322). These results suggest that immunosuppression in the testis may be restricting the ability of T cells to respond to SIV/HIV infection and allow virus persistence in this organ.
Overall, the immune-suppressive environment necessary to the testis homeostasis represents an Achilles heel when it comes to viral infections; activation of the acquired immune system is normally repressed in the testis, creating a shelter for viruses. In addition, inflammatory responses have negative consequences for the testis that can lead to sterility. In the human testis, weak antiviral innate responses, as observed following MuV and ZIKV infection in vitro and ex vivo, represent additional weaknesses that might promote viral persistence.
B. The Epididymis and Vas Deferens
1. Immune characteristics
Although autoantigenic spermatozoa accumulate in the epididymal ducts, transplantation experiments in rat epididymis and injection of spermatozoa in the stroma indicated limited immune protection compared with the testis (281, 317, 469). The epididymal epithelial barrier composed of apical tight junctional complexes is penetrable by macrophages, most lacking MHC class II (281, 463, 469), and by CD8+ T lymphocytes, both in elevated proportion in proximal regions of the epididymis (496, 497). DCs, present underneath and through the rodent epididymal epithelium (633, 637), are thought to be involved in the elimination of abnormal sperm cells and non-self-ascending pathogens (264). The fewer DCs present in normal human epididymis are located exclusively in the stroma (153). Immunoregulatory IDO, activin A, and TGF-β are highly expressed in rodent caput epididymis and may promote tolerance to sperm antigens (463).
2. Innate immune defenses
Epithelial cells from rodent epididymis express a wide range of PRR (TABLE 5). Activation of TLR3 and RIG-1 induced the expression of type I IFNs, ISGs, and pro-inflammatory cytokines, while DNA sensor ligand HSV-60 only induced the expression of IFNs and ISGs (781). Interestingly, an inverse gradient of expression of immunoregulatory genes (e.g., IDO and activin A) and PRR was observed along the length of the human epididymal duct (77, 78). This may lead to a tolerogenic-orientated environment in the regions proximal to the testis, and to a more vigorous antigen-specific immunity in the cauda, consistent with the need to protect sperm emerging from the testis without compromising the ability to respond to ascending infections. Rodent and human epididymis are major sites of production of β-defensins, a number of which are specific to this organ and display segment-selective expression along the length of the epididymis (321, 755). These small cationic peptides, essential for sperm maturation (768, 774), possess potent antimicrobial activity, including antiviral (291). The epididymis also secretes antimicrobial lysozyme and lactoferrin (275). In the vas deferens, TLRs 1–9 and 11 are constitutively expressed (527), and TLR2, 3, 4, and 9 agonists stimulated the secretion of CXCL1 (413).
C. The Seminal Vesicles and Prostate
1. Immune characteristics
The prostate is the male organ in which infectious and inflammatory occurrences are the most frequent, whereas pathologies of the seminal vesicles are rarely reported. Infections and inflammation also occur in this organ, however, but may be more silent. The continuous line of basal cells joined by tight junction at the basement of the prostatic epithelium, together with the expression of P‐glycoprotein at the apical pole of prostatic epithelial cells, contribute to restricting leukocyte passage and molecular movement through the prostatic epithelium (344). Like the seminal vesicles, the prostate is a strictly androgen-dependent organ. Testosterone possesses immunosuppressive properties and influences the innate immune responses as well as the outcome of infections and inflammation in the prostate (564). Thus, in rodents and other animals, castration is efficient at eliminating bacterial pathogens and dampening infection-related inflammation of the prostate (564). Little is known about the specific effects of androgens on host defense. In rat prostatic cells, testosterone negatively modulated the TLR4 pathway (563, 565). Also, testosterone maintained high levels of immunomodulatory factors in the prostate, such as galectin-1 (564). Nevertheless, the prostate is an immunocompetent organ. The numerous intraepithelial CD8+ T cells constitute a first line of defense against foreign agents reaching the prostate through retrograde flow (65). Even in the absence of pathology, most adult prostate tissues contain inflammatory infiltrates composed of T cells and macrophages (436, 507, 694).
2. Innate immune defenses
The prostate epithelial cells secrete an antimicrobial substance with potent antibacterial and antiviral actions such as defensins (563), protease inhibitors (e.g., secretory leukocyte protease inhibitor or SLPI) and collectin proteins (512, 513), while seminal vesicle epithelial cells secrete high levels of lactoferrin (736). Surprisingly, TLR 2–9 proteins were undetectable in human seminal vesicles (559), whereas mouse seminal vesicles cells in primary culture produced CXCL1 in response to TLR2, 3, 4, and 9 stimulation, although to a lower level than prostate and epididymis/vas deferens cells (413). Stromal cells isolated from patients with benign prostate hyperplasia (BHP) acted as antigen-presenting cells and responded to TLR agonists by producing pro-inflammatory cytokines/chemokines, except TLR9 (537). Mixed cell culture of normal rodent prostate responded to TLR2, 3, 4, and 9 agonists with the secretion of pro-inflammatory mediators and upregulation of TLR genes (215, 413, 414). These reports suggest that prostatic epithelial cells are equipped to resist infection, but their response to viral infections has never been studied. Additional studies are required regarding the immunity of primate seminal vesicles. This organ has been largely ignored and was recently found to seed virus and infected cells into semen (299).
D. The Penis
1. Immune characteristics
The penis is a major portal of entry for many pathogens and an immunologically active site. The immunity of the penis has been reviewed in detail elsewhere (558, 618), and only a few relevant key aspects are presented here. While the urethra displays classic mucosal effector features, the glans has even more activated natural killer (NK) cells and terminally differentiated effector CD8+ T cells (618). The penile urethra contains numerous polymeric immunoglobulin A (IgA) and IgM producing plasma cells (558). In addition, the urethral glands (Littré glands) secrete IgA that coats the urethral epithelial surface to form an immunological barrier against invading pathogens (558). Urethral swabs from HIV-highly exposed uninfected men contained HIV-1 specific IgA (86), suggesting effective responses that could be used for vaccination strategies. The penile foreskin is also able to mount a specific humoral response: exposing rhesus macaque foreskin to SIV induced SIV-specific IgG antibody and cytokine-secreting SIV-specific CD8+ T cells (588). The inner foreskin produces high levels of pro-inflammatory cytokines, a feature of inflamed epidermal barrier (385, 555) that is probably favored by the microbial community of the sub-preputial space (553).
2. Innate immune defenses
Both the foreskin and the urethral epithelial surface are coated with a mucus layer consisting of membrane-associated mucins, which in addition to lubrication play an important role in first-line immune defense by trapping microbes before they reach the epithelial surface (19, 596). Antimicrobial peptides are widely expressed all along the human penile mucosa: α defensin-5 is present in urethral secretions (551) and lysozyme in the urethral glands (Littré glands) (558) and foreskin (19); lactoferrin is expressed by urethral epithelial cells (558) and detected in the foreskin lamina propria (19); while SLPI is expressed abundantly by urethral epithelial cells and urethral glands (275, 517, 558). IFN-β is constitutively expressed in the foreskin epithelium and lamina propria, and by basal epithelial cells in the fossa navicularis (558). Highly localized IFN-α expression was noticed in the urethral epithelium of patients in the presence of high levels of TLR9 (558). Ex vivo infection of the inner foreskin with HIV increased the secretion of CCL5/RANTES, which mediated T cell recruitment into the epidermis, while decreasing the secretion of CCL20/macrophage inflammatory protein 3α (MIP-3α), enabling Langerhans cells to travel deeper into the tissue (776). HIV-1 infected CD4+ T cells formed viral synapses with foreskin keratinocytes and activated the MyD88-independent TLR4-NF-κB signaling pathway, leading to pro-inflammatory cytokine secretion, epidermal redistribution of Langerhans cells, and the formation of conjugates with T cells (777).
VI. CONSEQUENCES OF VIRAL INVASION OF THE MGT: FROM ANCIENT TO EMERGING INFECTIONS
A. Viruses as Enriching Colonizers: Benefits of Ancient Infections
Several viruses that infect somatic cells can persist lifelong in humans (e.g., EBV, HSV, HIV). However, the smartest way for a virus to permanently colonize a host is to integrate its DNA into the germline (i.e., the gametes and their progeny up to the early embryo) to be passed on vertically to future generations, invading every somatic and germ cell of the offspring. A number of viruses, mostly ancient but some contemporary ones such as HHV-6, have attempted this, leading to the integration of viral sequences into our ancestors’ genome which, if not detrimental, eventually became fixed (endogenous) in the population (183). This colonization of the germline has strongly shaped our evolution, adding new functions, and is still ongoing nowadays in mammals. Our review of the range of viruses, processes, and benefits associated with germline infection, along with elements suggesting initial male germ cell infection, appears below (FIGURE 8).

FIGURE 8.Integration of viruses in the germline and its consequences. A: pathway to endogenization and fixation of viral sequences in the human population and their beneficial effects. The integration of viral elements may nevertheless carry their burden of negative consequences: the reactivation of human endogenous retroelements (HERVs), when permitted by the innate and acquired immunity, has been incriminated in the development of inflammatory, autoimmune, and neoplastic diseases, although proof of causality is currently lacking (337, 338, 462, 717). B: estimated timeline of viral endogenizations in human ancestors. HHV-6, human herpesvirus-6.
1. Endogenous retroviruses
Sequencing of the human genome has provided proof of the enormous scale of viral infections that have led to the integration of viral genes into the chromosomes of the host germline, and transmitted vertically by Mendelian inheritance to become endogenous during evolution (258). Thus ~8% of the human genome consists of endogenous retroviral elements (ERVs), the remnants of ancient exogenous retroviruses that infected germ cells (258). More than 31 viral families of human ERVs (HERVs) have been characterized in the human genome (HERV K, W, H etc.), with over 100,000 copies of retroviral elements representing several amplifications (duplications, transpositions, etc.) and reinfection events during evolution (48, 340, 412, 525). Nearly all HERVs were integrated up to 100 million years ago (172), with the most recent integration episode (HERV K family members) estimated at between 100,000 and 1 million years ago (39, 737). ERVs have been found in all vertebrate genome sequences to date (48, 754), and endogenization of viral sequences is still ongoing, as shown in koalas (674).
a) male germline infection.
Several ERVs are located on the male Y chromosomes of human and NHPs (360, 638, 639), providing evidence of their transmission through the male gametes in these species. Specific characteristics of the Y chromosome, such as reduced recombination and low number of functional genes, may have prevented the loss of integrated sequences and allowed integration with no deleterious effect (360), thus favoring viral endogenization. While it is unknown whether male or female gametes or early embryos were preferentially targeted for initial infection, we speculate that in addition to Y chromosome integration, infection of the male germ cells might offer a selective advantage for viral dissemination compared with infection of the female gametes. Unlike female germ cells, the male stem germline continues to divide in adulthood and gives rise to an indefinite number of sperm cells, allowing numerous vertical transmissions of integrated sequences with endogenization potential.
b) ancient and contemporary lentiviruses.
While integration into the germline is commonplace in many retroviral genera and hosts, it was thought until relatively recently that lentiviruses (the group of retroviruses to which HIV belongs) were entirely exogenous. This was proven wrong by the discovery of ancient endogenous lentiviruses in different mammals (231, 232, 341, 346). The finding of endogenous prosimian SIV in lemurs raises the possibility that contemporary HIV and its simian counterpart SIV might one day also become endogenous. The detection of HIV DNA within isolated testicular germ cells and spermatozoa from patients (691, 722) has been controverted. Interestingly, results from our laboratory show that HIV can bind, enter, and integrate its genome into the male germ cell genome in vitro, albeit inefficiently (unpublished data). Although primates have evolved a variety of cellular factors to block viral infection, viruses evolve much more quickly than their host and can counteract these innate immune factors (155). The repertoire of antiviral factors expressed by human male germ cells, and in turn their ability to block integration of contemporary viruses, is currently unknown.
c) host virus interaction and added functions.
ERVs are usually inactivated by genetic mechanisms [e.g., deletions, inversions, or point mutations in the open reading frames of viral proteins under the pressure of host innate factors such as apolipoprotein B mRNA editing enzyme catalytic polypeptide-like (APOBEC)] and epigenetic mechanisms (323) and therefore cannot produce infectious viral particles, with a few exceptions (592). Interestingly, it was recently demonstrated in the koala that older endogenous retroelements recombine and degrade the new colonizing retrovirus genome, thus contributing to their inactivation (401). Nevertheless, several integrated retroviral DNA copies are still transcribed and encode original or truncated viral proteins (e.g., HERV-K copies during embryogenesis) which can add functions to the host. In addition, the ability of ERVs to recombine and transpose into the host genome and to modify the nearby gene expression through their own sequences [e.g., the long terminal repeat (LTR) promoter and enhancer element in retroviruses] is an important source of genomic and regulatory variability (338). Understandably, the activity of ERVs is under tight complex regulation by the host, especially in the germ cells and during early embryo development, where ERVs are sequentially and specifically expressed or silenced to protect them from excessive re-infection or transposition (592).
I) Role in reproduction. The best known beneficial impact of retrovirus insertions identified to date is on the acquisition of viviparity through the development of effective placentation. Retrovirus envelope genes encode glycoproteins with fusogenic properties, a function necessary for virus entry. The envelope protein from HERV-W (endogenized in our primate ancestors 25 million years ago) was purloined to drive cell fusion in the placenta to form the syncytiotrophoblast, and was hence renamed syncytin-1. Syncytin-2, an envelope protein highly expressed in the placenta, also derives from an ERV (HERV-FRD) (160, 270). A similar process has occurred in various mammals (159), suggesting a nonrandom pattern of co-option (i.e., usage of a function that is not the original one). The syncitiotrophoblast is essential not only for invasive placental development but also for prevention of immune rejection of the fetus. Interestingly, syncytin-2 also suppresses immune recognition and could thus contribute to the mother’s immune tolerance of the fetus (160, 308). A role for syncytins in the cell fusion of osteoclasts (648) and muscle cells (570) was also evidenced. Other ERV elements play immunoregulatory roles in the placenta. For instance, the tissue-specific expression in human trophoblasts of HLA-G, central to immune tolerance during pregnancy through the inhibition of NK cell mediated cytotoxicity, is controlled by the LTR of human ERV1 (182). Further critical roles for HERVs in human reproduction were recently discovered. A number of converging elements indicate that HERVs actively contribute to the maintenance of pluripotency in early embryos (462). In male germ cells, the strong promotor activity of the ERV9 LTR controls the expression of unique isoforms of p63 that preserve genetic integrity by suppressing cell proliferation and inducing apoptosis upon DNA damage (52). The exclusive expression in human testis tissue of a number of HERV transcripts has been reported (639), suggesting other roles for HERVs.
Overall, several of the ERVs which colonized the germline of our ancestors can be considered as the initiators (e.g., role in placentation) and/or the guardians (e.g., role in maintaining genetic integrity in germ cells and pluripotency in the embryo) of human reproduction.
II) Protective role against infections and pathological processes. The impact of ERVs on humans extends beyond reproductive functions. One well-identified positive effect of ERVs is the induction of direct protection against exogenous viruses by interacting with different steps of their life cycle, a “fighting fire with fire” process. For instance, competitive binding to the cellular receptor mediating viral entry can occur between the envelope proteins of exogenous virus particles and that encoded by ERVs, as found in mice, sheep, and cats (32, 421). Among other protection mechanisms is the association, during assembly, of particles newly released from the exogenous virus of defective viral components encoded by the ERVs, a phenomenon observed for JRSV in sheep (32) and suggested for HERV-K gag protein during normal embryogenesis (262). Finally, ERVs can boost innate and acquired immune functions. Only a couple of examples are presented here (for a complete review, see Ref. 338). A number of ERV LTRs contain interferon-stimulated response elements (ISREs) which, after IFN stimulation, can enhance adjacent genes critical for antiviral and pro-inflammatory responses. The HERV-K protein Rec can inhibit exogenous viral infections in human pluripotent cells by increasing the innate antiviral response, and it is suspected that it plays an immune-protective role during early embryonic development. In transformed cells, the expression of endogenous retroviral elements is no longer silenced and can induce innate and acquired immunity to target nonhealthy, ERV-producing cells, facilitating their elimination. Endogenous retrovirus proteins are part of the so-called “cancer testis antigens,” a set of “self” proteins not expressed in healthy cells other than testicular germ cells, where they are protected from immune recognition by the immunosuppressive testis environment.
In summary, ERVs greatly contributed a range of functions in their hosts. However, in specific conditions, they may be implicated in the development of human diseases. Indeed, when permitted by the immune system, the reactivation of HERV has been associated with a plethora of syndromes affecting a wide range of organs. These include solid and blood tumors (417), neurological disorders (e.g., schizophrenia, autism) (374, 468), and autoimmune diseases (e.g., rheumatoid arthritis, systemic lupus erythematosus, multiple sclerosis) (249). In both human prostate and testicular cancers, epigenetically driven differential and specific expression patterns of the HERV-K (HML-2) loci were observed and associated with cancer establishment (152, 241, 242, 278, 610, 658). Nevertheless, formal proof of ERVs implication in diseases causality is currently lacking.
2. Germline integration of non-retroviral viruses
Additional endogenous viral elements that do not belong to the retrovirus family have been discovered in the genome of a variety of organisms (6, 292, 339). In vertebrates, a range of DNA viruses is involved, including Herpesviridae (in humans with HHV-6 and in prosimian Tarsier), Parvoviridae (in eutherians, marsupials, and birds), Circoviridae (in eutherians, marsupials, and amphibians), and Hepadnaviridae (in birds). Two families of RNA virus have also been endogenized: Bornaviridae (in mammals including humans and in marsupials) and Filoviridae (in marsupials and several mammals, such as bats) (6). In fact, it is becoming increasingly clear that virtually any type of virus can be endogenized.
a) mechanisms of integration.
The mechanisms by which non-retroviral viruses integrated the germline and became endogenous are not yet fully understood. Albeit less often than with retroviruses, some DNA viruses can undergo integration through homologous mechanisms [e.g., herpesviruses (480)] or nonhomologous mechanisms [e.g., hepadnaviruses, adeno-associated viruses which are parvoviridae dependovirus (333)] of recombination. As for non-retroviral RNA viruses, these need to undergo processes that are unusual and do not normally occur in their life cycle, such as reverse transcription of viral RNA to DNA, entry into the nucleus and integration. The integration of DNA sequences complementary to non-retroviral RNA viruses into the DNA of in vitro infected mammalian cells has been reported (362, 772). Interestingly, in relation to the infection of human male germ cells with ZIKV (441, 581), an arbovirus of the Flaviviridae family, chromosome integration was reported for another arbovirus (Sindbis virus) (773) and for animal Flaviviridae HCV (636). The mechanisms for such integrations could involve endogenous reverse-transcriptase activity, revealing interaction between endogenous retroelements and exogenous RNA viruses (362).
b) elements in favor of viral integration into male gametes.
Epididymal spermatozoa may represent a preferential target for the transmission of non-retroviral viral sequences. Spadafora and co-workers (234) demonstrated that mammalian spermatozoa can take up foreign DNA and RNA in vitro, transfer them into their nuclei, and reverse-transcribe RNA into cDNA fragments, due to the endogenous reverse-transcriptase activity stored in sperm nucleus and encoded by LINE-1 retrotransposons. The capture of foreign nucleic acids is inhibited by seminal fluid and may therefore not happen in ejaculated spermatozoa. Endogenous reverse-transcription expression, normally repressed in differentiated nonpathological tissues, is increased in germ cells and early embryos (615). Reverse-transcription and integration of poliovirus RNA into the sperm nucleus was demonstrated (234). The authors speculated that the viral cDNAs produced were amplified by a DNA-dependent RNA polymerase, released from the spermatozoa, and taken up again by further spermatozoa, thus spreading foreign nucleic acids among the sperm (653). In mouse models, foreign nucleic acids within spermatozoa nuclei were transmitted to oocytes as low-copy number, extrachromosomal sequences transcriptionally competent, and were mosaic propagated in tissues from the offspring, introducing new genetic traits in a non-Mendelian fashion (653). Alternatively, early embryos could support integration of viral cDNA sequences, leading to their endogenization. Interestingly, fragments of foreign DNA were integrated in the mouse sperm genome following recombination with chromosomal DNA at preferential sites (418, 783), suggesting another potential endogenization mechanism for non-retroviral viruses. However, a sperm endonuclease cleaves foreign DNA, and when potently activated degrades the host DNA, leading to cell death (420), probably controlling the integration processes.
Non-retroviral viruses endogenized in humans or with potential for germline integration are presented in TABLE 6, along with demonstrated or putative effects.
Viral Family | Endogenous Sequences in | Postulated Role | Potential for Integration in Human |
---|---|---|---|
Filoviridae (RNA virus) | Bat, marsupials (49) | Overexpression of a viral glycoprotein may prevent EBOV disease (424) | EBOV and MARV in testis of patients and NHPs (111, 117, 430, 541, 766) |
Bornaviridae (RNA virus) | Human and other vertebrates (49, 297), marsupials | Protect against exogenous related virus (Borna disease virus) in ground squirrel through an endogenous viral protein expression and incorporation in exo-virus (197, 295); integration in piwi-interacting RNA (small noncoding RNA expressed in germ cells and early embryo) to silence transposons (reviewed in Ref. 295); silencing of related exogenous viral RNAs (295) | |
Parvoviridae: AAV (DNA virus) | Parvovirus-related DNA sequence in human genome (single integration more than 98 million years) (397) | Establish site-specific integration into human chromosome and latency (369, 601); detection in testis tissue (454) | |
Hepadnaviridae: HBV (DNA virus) | Zebra finch | Benefits: in utero acquisition of HBV in human enhances newborn immune cell ability to respond to unrelated pathogen exposure through a “trained immunity” process (296, 515); Deleterious consequence: chronicity | HBV integrates into host genome; integrates into sperm from patients (269, 304); spermatozoa with integrated HBV can fertilize oocytes in vitro (304) but viral gene expression from sperm-introduced HBV may interfere with embryonic development and cause abortion (12, 366) |
Herpesviridae: HHV-6 (DNA virus) | Human (1% of the population) (480), endogenization date back 14,000–35,000 yr for some individuals (769), ongoing integration in germline (480)? | Full set of intact viral genes with potential to reactivate in patients: impact on health (769)? | HHV-6 integrates into telomeres of chromosomes (26); vertical transmission of chromosomally integrated HHV-6 from father’s spermatozoa and from mother (26, 130, 273, 479, 672, 725) |
B. Viruses as Aggressors: Deleterious Effect of Contemporary Viruses on the MGT and Offspring
FIGURE 9 details the negative effects on the MGT of a range of viruses.

FIGURE 9.Deleterious consequences of viral infections of the human male genital tract (MGT) at the individual (A) and population and offspring levels (B). Examples of viruses associated with these consequences are shown. A: viral infections of the MGT may have transient or prolonged negative effects on semen and sperm parameters, and induce infertility and endocrine disturbances linked to modifications of testicular hormone secretion. These alterations can derive from a direct or indirect effect of the infection on the testis (see FIGURE 10) or from the infection and inflammation of the accessory glands, leading to impaired sperm maturation, gametotoxic effect, or ductal obstruction. Spermatozoa and their testicular germ cell progenitors can be targeted by viruses, thus presenting a risk for vertical transmission. Viral invasion of male genital organs also increases the risk of tumorigenesis (e.g., penile cancer) that might be initiated through complex virus-host interactions including inflammatory mechanisms. Viral etiology in prostate and testicular cancers, however, is currently unproven. The MGT can allow viral persistence and constitute a pharmacological sanctuary in patients under therapy. B: the presence and persistence of virus in the MGT may lead to viral excretion in semen, contributing to horizontal viral dissemination and indirect vertical transmission, the latter potentially inducing congenital disorders such as miscarriage and fetal malformation [e.g., Zika virus (ZIKV), cytomegalovirus (CMV)]. Direct infection of the embryo may derive from infected or virus-bound fertilizing spermatozoa. Successful germline infection with viral integration in our ancestors has led to viral sequence transmission through gametes and endogenization at the population level [e.g., human endogenous retroelements (HERVs), human herpesvirus-6 (HHV-6)]. In the case of HHV-6, it is unclear whether this process is still ongoing. Damaging effects of viral infections on sperm and natural barriers protecting the oocytes might restrict such transmission, only reported in vitro. HVB, hepatitis B virus; HCV, hepatitis C virus; HSV, herpes simplex virus; HPV, human papillomavirus; MuV, mumps virus; EBOV, Ebola virus.
1. Viral infections and infertility
Male infertility currently affects 20–50% of couples seeking medical assistance for procreation, of which 6–15% are attributable to infections (185, 624). Infertility is defined as the failure to achieve a clinical pregnancy after 12 mo or more of regular unprotected sexual intercourse (765). In addition to identified factors such as varicocele, cryptorchidism, as well as endocrine and genetic disorders, male infertility is considered idiopathic (i.e., of unknown origin) in a large number of cases. The chemical and biological environment that includes pathogen exposure is increasingly acknowledged as a putative cause of infertility (233, 616, 728). Following viral infection, sperm parameters (e.g., sperm count, motility, number of abnormal spermatozoa, etc.) and/or semen parameters (e.g., volume of seminal plasma, viscosity, pH, enzyme concentrations, etc.) can be transiently or, more rarely, permanently altered to various degrees (see below). In cases of severe defects, this may lead to infertility. While there is no established semen parameter cut-off to predict loss of fertility, apart from azoospermia (absence of spermatozoa in semen), infertility is commonly associated with sperm alterations such as oligozoospermia (low sperm count below 39 × 106 spermatozoa per ejaculation or concentration below 15 × 106 sperm per ml), asthenospermia (percentage of progressively motile sperm below 32%), or teratospermia (percentage of morphologically normal sperm below 4%) (387).
In the context of viral infections, alterations of sperm and semen parameters may result from distinct nonexclusive phenomena:
MGT tissue inflammation. While defending against pathogens is critical for controlling invasion and preventing virus-induced damage, high levels of inflammation in the testis (orchitis), epididymis (epididymitis), and prostate (prostatitis) can be deleterious for reproductive functions (16, 185, 713). Orchitis is most commonly induced by a viral infection and usually involves one testicle only (613). Histologically, orchitis is characterized by multi-focal immune cell infiltrates composed of granulocytes, macrophages, and lymphocytes, localized in the interstitial tissue and inside the seminiferous tubules. Orchitis is associated with thickening of the lamina propria bordering the tubules (i.e., peritubular cells and extracellular matrix) and degeneration of the germinal epithelium (185). Dysregulation of the testicular cytokinetic microenvironment disrupts both steroidogenesis (271, 697, 752) and spermatogenesis (53, 325, 602, 681). For instance, during orchitis, infiltrating macrophages and mast cells produce TNF-α that induces germ cell apoptosis (681), modifies peritubular cell secretions, and induces fibrosis (442). Inflammation of the epididymis and prostate, more commonly induced by bacterial infections but potentially also by viral infections, can also alter sperm parameters or cause irreversible ductal obstruction and fibrotic tissue remodeling, and account for up to 12% of infertility cases (150, 185, 463, 713).
Viral replication within the cells and organs of the MGT may alter their function and integrity, through altered endocrine, paracrine, or other biochemical secretions; breakage of the BTB (favoring the induction and leakage of sperm antibodies in the lumen of the MGT and germ cell infection); gametotoxic effect; sperm DNA damage; epigenetic modifications, etc.
A systemic effect of the disease. Febrile episodes in acute infections can alter fertility because the elevated testis temperature impairs spermatogenesis, a temperature of 34°C being necessary to preserve testis homeostasis. In chronic infections, increased levels of oxidative stress, to which spermatozoa are especially sensitive, may damage sperm function. Modifications of the release of the luteinizing hormone (LH) and folliculo-stimulating hormone (FSH) from the pituitary, which control Leydig cell and Sertoli cell endocrine functions, respectively, can decrease sperm counts by affecting the secretion of testosterone and inhibin B, which are necessary for spermatogenesis.
The viral infections, classified as either acute or chronic, that have been associated with altered sperm/semen parameters or male infertility and their postulated mechanisms of action are presented below and in FIGURE 10.

FIGURE 10.Mechanisms for testis function perturbation caused by viral infections. Viral infections may impact testicular endocrine and exocrine functions through several non-mutually-exclusive mechanisms. A: viral infections can indirectly perturb testis functions through a systemic effect on the hypothalamic-pituitary-testicular modifying testicular hormone production [e.g., human immunodeficiency virus (HIV)/AIDS] and through fever-induced elevation of testis temperature, which is deleterious for germ cell differentiation (e.g., influenza, chicken pox, pneumonia). B: testis inflammation (orchitis) can occur following a viral attack of the testis [e.g., mumps virus (MuV)] or after systemic infection/generalized inflammation (e.g., influenza virus). This orchitis can transiently or permanently alter testis morphology and functions, leading to sterility if severe and bilateral. C: viral replication in testicular cell types can damage their function, impair immune-privilege, and disrupt the complex intercellular and paracrine cross-talks required for steroidogenesis and spermatogenesis [e.g., MuV, Marburg virus (MARV)]. Infection or viral attachment to germ cells and spermatozoa can also result in horizontal and vertical transmission (e.g., hepatitis B virus, HIV, human papillomavirus). Furthermore, testicular cells may constitute a reservoir for viruses (e.g., Zika virus, Ebola virus, and MARV). GnRH, gonadotropin releasing hormone; FSH, follicle stimulating hormone; LH, luteinizing hormone; SC, Sertoli cell.
a) altered sperm parameters/infertility following acute viral infection.
I) Mumps virus. MuV is a well-established etiological agent of orchitis, which occurs in ~10–20% of infected post-pubertal men (437). Consecutive testis atrophy happens in ~50% of these cases and can be associated with low sperm count (oligospermia) and hypofertility, while bilateral atrophy leading to sterility is rare (594). The histology of the testis following MuV infection has been described in detail (138). MuV was rescued from the testis of infected patients (59), but its testicular target cells and the mechanisms underlying germ cell degeneration remain elusive. MuV replicated in human Leydig cells in vitro and decreased their testosterone production (244), in line with decreased testosteronemia in patients (4, 7). In a mouse model, MuV also infected primary Leydig cells, as well as Sertoli cells and testicular macrophages, but did not target testicular germ cells (748). Several nonexclusive hypotheses could explain the MuV-induced impaired spermatogenesis in humans: 1) change in testis temperature due to high fever; 2) congestion of the seminiferous tubules caused by the interstitial edema (425); 3) decreased testosterone production by infected Leydig cells (244); 4) alteration of the paracrine control of spermatogenesis exerted by somatic cells (Sertoli cells, Leydig cells, testicular macrophages) as a result of their infection; and 5) germ cell apoptosis triggered by CXCL10 production by infected Sertoli cells, as observed in the mouse (325).
II) Zika virus. ZIKV infection in men was associated with decreased sperm count and increased sperm abnormalities up to 90 days after the onset of symptoms, along with prolonged viral excretion in semen and the association of infectious viral particles with spermatozoa (307, 326). Relatively mild endocrine perturbations were observed in the early days post-symptoms onset (e.g., lower inhibin β concentration), without any significant modification of testosterone levels (326). The sperm parameters recovered after 4 mo (326). Hematospermia (blood in semen, indicative of accessory gland infection/inflammation) and an increased leukocyte count occasionally occurred (307). These anomalies were observed irrespective of virus detection in the semen (307), suggesting either a systemic effect independent of MGT infection or persistence of the damage after viral clearance. We revealed that ZIKV efficiently replicates in human testis explants and targets a wide range of testicular cells including germ cells, which were found to be exfoliated in the semen from ZIKV-positive men (441). No cytopathic effect was observed and the general testis structure was preserved, including tight junction protein ZO-1 expression by Sertoli cells. Basal testosterone and inhibin B levels were unchanged by the infection. Altered sperm parameters could result from 1) ZIKV replication in Sertoli cells, Leydig cells, and testicular macrophages, altering the tight paracrine control of spermatogenesis and consequently the sperm count; 2) ZIKV replication in peritubular cells, the contractile ability of which is essential for the release of spermatozoa in the testis lumen; 3) ZIKV infection of immature germ cells, which consequently alters their differentiation; and 4) association of ZIKV with late germ cells/spermatozoa. Semen alterations in ZIKV-infected men could also result from infection of the epididymis/accessory glands or a systemic effect (e.g., fever). Immunodeficient mouse models [e.g., knockout mice for functional type I IFN receptor (IFNAR−/−) or treated with anti-type I IFN antibodies] were developed to study ZIKV infection (483, 657). Wild-type mouse strains are normally resistant to the virus, the replication of which is efficiently inhibited by murine type I IFN, whereas in humans, type I IFN signaling is counteracted by ZIKV protein NS5, allowing viral replication. A high level of testis infection was described in mouse models (96, 110, 151, 247, 257, 410, 441, 702) leading to large inflammatory infiltrates and consequently testis atrophy, collapsed testicular hormone levels, breakdown of the BTB, germ cell degeneration, and infertility. In addition to testis, high levels of ZIKV replication were consistently detected in mouse epididymis (96, 110, 151, 157, 247, 257, 410, 702), where the virus associated with epithelial cells and/or spermatozoa. Mouse prostate and seminal vesicle infections were not systematic (110, 157, 410, 445). In ZIKV-infected NHPs (287, 364, 390, 524), the virus was occasionally detected in the testis after day 7–8 post-infection (364, 390, 524), without tissue alteration, and scarcely infected cells were observed in the prostate and seminal vesicles (524), along with low levels of viral RNA similar to other non-MGT tissues (287). In three of six infected baboons, a marked transient decrease in total and motile sperm count was observed, but these differences did not reach significance (154). Overall, the drastic effect of ZIKV on the MGT in mice significantly differs from that in ZIKV-infected men and NHP, in whom orchitis has never been reported and modifications of sperm parameters appear milder and transient. Higher levels of viral replication and inflammation in immuno-deficient mice are probably responsible for this discrepancy.
III) Other viruses. A number of systemic acute infections with viruses never detected in semen or MGT organs trigger transient modifications of sperm parameters. Decreased sperm count, motility, and altered morphology occurred 8–11 wk after febrile episodes caused by influenza virus (619), chicken pox, or pneumonia (415). Sperm DNA integrity was compromised after influenza virus infection (175, 619). In the absence of orchitis and viral material detection in the MGT, transiently altered sperm parameters could result from high fever-induced damage of spermatogenesis, similar to that evidenced after sauna exposure (76). Disturbances of the male endocrine system caused by modifications of the hypothalomo-pituitary-gonadic axis could also be involved.
b) altered sperm parameters or infertility associated with chronic viral infections.
I) Hepatitis viruses. A study using a large national insurance data set from the Taiwan health service (5,138 men with HBV vs. 25,690 uninfected men) found a significant increase in the incidence of infertility associated with chronic HBV infection, after adjusting for a number of covariates (666). Because the majority of HBV infections in Taiwan occur during the neonatal period, the individuals studied had a longer exposure time to HBV than Western populations, who mostly acquire HBV during adulthood. However, several Western and Eastern studies have reported altered sperm parameters and a negative impact on fertilization in men chronically infected with HBV, including decreased sperm count and/or mobility, increased apoptosis, aneuploidy, and/or number of abnormal spermatozoa (382, 402, 477, 516). Disease-induced oxidative stress, to which spermatozoa are especially sensitive, is evident in semen from HBV-infected individuals (562). In vitro, viral protein S was found to induce sperm mitochondrial dysfunction (331). HBV DNA integrated into the sperm chromosomes of patients (132, 269, 304) and induced chromosomic instability and DNA damage (304, 477).
Chronic infection with HCV has also been linked to modification of semen parameters. Lower sperm mobility and morphology as well as decreased serum levels of inhibin B and testosterone were described in several studies in HCV-infected versus uninfected men (161, 288, 289a, 402, 477, 712), except one (213). Antiviral treatment with IFN and ribavirin improved the hormonal pattern but further degraded sperm quality (161, 288). Semen viremia did not correlate with altered semen parameters, suggesting that HCV does not exert a direct negative effect on sperm (67). HCV in the male partner had no negative impact on pregnancy outcome during assisted reproduction procedures (67, 757).
A strikingly high prevalence of HEV in semen from 185 infertile Chinese men was recently reported (28 vs. 0.5% in general population) (303). Oligospermia and asthenospermia (i.e., >65% of immobile spermatozoa) were observed in HEV+ men. Experimental HEV infection of rhesus macaques triggered leukocyte infiltration in the epididymis and testis, damage to the seminiferous tubules, and decreased testosterone levels (303). These intriguing data demand further investigations.
II) Human immunodeficiency virus. In the late stage of HIV infection (AIDS), oligospermia or azoospermia has been frequently reported, together with orchitis and severely damaged testis morphology with immune cell infiltration (138). Low testosterone levels (which contribute to cachexia) were encountered with AIDS, along with normal or elevated LH and FSH levels, implying primary testicular failure (135). Rather than HIV infection of testicular cells, the systemic debilitating effect of the disease together with generalized inflammation most probably accounts for the orchitis and hypogonadism in AIDS patients. Thus, although HIV-1/SIV infects testicular macrophages and T lymphocytes early on in the primary infection throughout the course of the disease (691), testicular morphology was preserved in asymptomatic, chronically infected patients (486), as well as in acutely or chronically infected macaques (693). No inflammatory infiltrates or elevated cytokine levels were apparent in infected men or macaque testes at the asymptomatic chronic stage (486, 693), unlike in AIDS-deceased patients and macaques (465, 557). Testosterone-producing Leydig cells were not infected in vivo or in vitro (465, 486, 557, 590, 625, 693, 738). Lymphocytic infiltrations of the testis and interstitial fibrosis in AIDS patients were associated with a decrease in Leydig cell number, which might partly account for the decrease in testosterone (135). IL-1, a pro-inflammatory cytokine released by phagocytic cells, inhibited the human chorionic gonadotropin (hCG)-stimulated steroidogenesis in Leydig cells in vitro (83), adding another potential cause for testosterone decrease. The face of HIV disease has dramatically changed with the introduction of combined antiretroviral therapies. The efficient control of blood viral load achieved by lifelong antiretroviral treatments has massively improved the quality of life of infected individuals, allowing an almost normal life span. Most of the 37 million HIV-infected individuals are of reproductive age (745). In many developed countries, HIV-infected men with controlled viral loads have been offered medically assisted reproduction to avoid transmission to the partner and embryo. Although still controversial (353), natural conception is now additionally proposed in some countries as a relatively safe/low risk alternative when specific conditions are met (e.g., undetectable viral loads in blood and semen for over 6 mo) (708). Semen parameters in HIV+ men attending fertility clinics have been extensively studied. Overall, relatively mild alterations were reported, along with modifications to the sperm mitochondrial DNA (324, 691). These alterations, observed in asymptomatic men with controlled viral loads, are believed to occur as a secondary effect of antiretroviral molecules and not as a consequence of the infection itself. Many studies have pointed at a deleterious effect of antiretrovirals (194, 545, 691), and an eloquent demonstration was provided with the longitudinal follow-up of a cohort of men before and after treatment initiation, showing a degradation of sperm parameters during treatment (384).
III) Human papillomaviruses. HPV positivity in semen was recently associated with a twofold increased risk of infertility in a meta-analysis of 31 studies comprising 2,122 men from the general population (HPV prevalence of 11.4%) and 3,072 men attending fertility clinics (HPV prevalence of 20.4%) (409). High-risk HPV16 was the most common type in semen in both populations, followed by high-risk HPV56. A previous meta-analysis reported a lower but still elevated HPV prevalence of 10% in the general population versus 16% in men seeking fertility treatment (379). HPV in semen has been associated with decreased sperm motility in some studies (total of 1,944 participants), but not in others (total of 1,143 participants) (191, 408), and with the presence of anti-sperm antibodies, which may interfere with sperm motility and oocyte binding (211). In addition to its presence in exfoliated epithelial cells in semen, HPV genome was detected on the sperm surface in infertile patients (191). An in vitro study described increased DNA fragmentation in sperm transfected with HPV DNA, but these findings were not substantiated in studies on semen from HPV-positive versus HPV-negative fertile patients (118, 335). Lately, HPV vaccination in men with HPV in semen has reduced virus prevalence, improved sperm motility, and decreased anti-sperm antibodies, probably through immunity stimulation, while improving pregnancy and live birth rate (190, 212).
IV) Herpes viruses. HSV-2 has been long suspected of being involved in infertility (135), but proof is still lacking. Several studies across the globe have reported a correlation between HSV-1 and -2 in semen and altered sperm and semen parameters, including low sperm count and motility, as well as reduced seminal volume and concentrations of epididymal (neutral α-glucosidase) and prostatic (citrate) molecules (53, 334, 373, 471). Together with hematospermia, these biochemical modifications suggest an impact of HSV infection on the epididymis and accessory glands physiology (53, 373). Two studies failed to substantiate the effect of HSV infection on sperm parameters, but their conclusions could have been impaired by the very low number of HSV+ patients tested (n = 4) (54, 503) and by assessment of HSV infection through antibodies rather than virus detection (478). HSV-2 bound to human ejaculated sperm in vitro and slightly decreased their motility, but HSV-2 binding was impaired by seminal plasma (529). HSV-1 and -2 DNA were reported in spermatogenetic cells from men and guinea pigs using in situ hybridization (256, 370). Since all of these studies only investigated infertile men or men seeking fertility evaluation, it would be important to compare the prevalence of HSV-1 and HSV-2 in semen from fertile versus infertile men in large cohorts with similar age, race, geographical, and socio-economic criteria (case-control study) to confront the conflicting data on the potential involvement of HSV in male infertility.
The impact of other Herpesviridae (CMV, HHV-6, and EBV) on sperm or semen parameters has been investigated, and no significant or consistent association has been demonstrated (54, 163, 334, 499, 500, 503, 758).
2. Viral infections and cancers of the MGT
Viruses can initiate or favor carcinogenesis through complex interplay with the host genetics, immunity, and inflammation (461), and seven viruses are well recognized as having a transforming or tumorigenic activity in humans (EBV, KSHV, HPV, HBV, HCV, HTLV, and Merkel cells polyoma virus or MCPyV) (60).
a) penile cancer.
Penile cancer accounts for 0.5% of all cancers in men and generally develops after the age of 60, with incidence estimated to be between 0.2 and 2.2 in 100,000 per year (352). High-risk HPVs, mainly HPV-16 followed by HPV-18 and HPV-45, are established risk factors in squamous cell carcinoma of the penis (284).
b) prostate cancer.
Prostate cancer is the most commonly diagnosed cancer in men worldwide (1.4 million cases in 2016) and the fifth most common cause of cancer death (240). Established risk factors for all prostate cancers include age, race/ethnicity, family history, and genetic variants, while lifestyle may influence cancer aggressiveness (540). Chronic prostate inflammation (that could be induced by a viral infection) is thought to increase the risk of prostate cancer (436), and a link between prostate cancer, sexual activity, and sexually transmitted infection was reported in some studies (306). However, despite years of investigations on HPV (3, 35, 301, 675, 756, 760), KSHV, EBV, BK virus, CMV (3, 301), and the “rumor” virus xenotropic murine leukemia related virus (XMRV) (531) through serology testing or prostate tissue analysis by PCR, in situ hybridization, or immunohistochemistry, there is no compelling evidence for a viral etiology of prostate cancer. Nevertheless, a few studies suggested that KSHV and HPV may represent a risk factor for prostate cancer aggressiveness (239, 494, 756). A recent metagenomics study highlighted two viruses not investigated so far, MCV and JC polyomavirus (JCPyV), in a small number of prostate cancer cases (642).
c) testis cancer.
Testicular cancer generally affects men between the ages of 15 and 35 and accounts for ~1% of male cancers (561). Testicular cancer is considered to primarily stems from developmental abnormalities during fetal life (641), with exposure to various factors during adolescence and adulthood potentially promoting its development (446). While there is presently no convincing evidence to support an association of any of the viruses tested (EBV, HPV, CMV, parvovirus B19, HIV) with testicular cancer (284, 762), a significant link between testicular cancer and a history of epididymo-orchitis was recently evidenced (332). This finding is in support of a role for pathogen-induced inflammation.
3. Vertical transmission of viruses
Viruses present in semen may indirectly contaminate the offspring by infecting the female partner during conception and pregnancy, or be directly transmitted to the oocyte and embryos, for instance, through infected spermatozoon.
a) indirect transmission through semen.
Among the pathogens that induce congenital disorders at various time points during pregnancy (originally grouped under the acronym TORCH for Toxoplasma, Others, Rubella, CMV and HSV, with an expanding list of “others”) (97, 121, 283) are a number of viruses present in semen (HSV, CMV, ZIKV, parvovirus B19), all sexually transmissible to women, except parvovirus B19. Other viruses transmissible to women through semen, such as HIV and HBV, are vertically transmitted without inducing congenital defects. In utero infection by HBV represents a predisposition to chronicity, possibly due to fetal immune tolerance to foreign antigens (51, 296, 470). Interestingly, however, HBV infection during fetal life enhances the ability of the newborn immune cells to respond to unrelated pathogen exposure through a “trained immunity” process (296, 515). A high viral load in the mother is a significant risk factor for vertical transmission (635), which may result, among other routes of infection, from exposure of the mother to infected semen, followed by hematogenous spread and transplacental infection. Importantly, enhanced in utero transmission of ZIKV was observed in a mouse model following exposure to infected semen when compared with subcutaneous infection mimicking mosquitoes bite or with intra-vaginal inoculation (156). Sexual transmission increased ZIKV dissemination in the female genital tract and led to ovary infection. Surprisingly, intravaginal inoculation of ZIKV did not enhance fetus infection as sexual transmission did, which might be due to differences in virus dose, post-coitum semen deposition directly into the uterus in mice, or semen-induced inflammation in the female tract. Thus, in addition to transplacental hematogenous spread, ascending viral infection of the female tract following semen deposition might constitute a particularly efficient vertical transmission route that deserves further investigation.
b) direct transmission through sperm.
Viruses that infect or associate with spermatozoa [e.g., HBV (132, 269, 304), ZIKV (326, 427), HPV (191), HSV (370, 529), HIV (298), CMV (499)] have the potential to be transferred to the embryo upon fertilization. While there is no direct evidence of sperm-mediated vertical transmission of nonendogenous viruses in humans, human spermatozoa naturally or experimentally infected with HIV, HBV, and HPV transmitted these viruses to hamster zona-free oocytes and early embryo following in vitro fertilization (11, 192, 722). However, HBV DNA integration into the sperm genome of patients triggered chromosome instability and DNA damage (304, 477). Spermatozoa transfected with HBV were prone to apoptosis and had reduced fertilization capacity when using human oocytes (305). HIV became attached to ejaculated spermatozoa in vitro, essentially through heparan sulfate proteoglycans (HSPGs), but could not enter these cells (88, 691). HIV virions bound to the surface of the sperm therefore have to cross the barriers of the zona pellucida and the membrane of the oocytes during natural conception. A meta-analysis covering 11,585 cycles of assisted reproduction among 3,994 women with HIV+ partners showed that sperm-mediated HIV transmission never occurred following sperm washing (763). Although HIV DNA was detected in patients’ sperm chromosomes (722), and the data from our laboratory suggest that HIV can enter and integrate into the genome of testicular germ cell in vitro, this event is rare, making successful vertical transmission a very low probability. HPV binds to sperm head in patients, using HSPGs as an attachment receptor, as with HIV-1 and HSV (191, 192). On the basis of in vitro experiments and reports in HPV+ patients, it has been postulated that HPV carried by spermatozoa can be transferred to fertilized oocytes and impair embryo development into blastocysts and the invasiveness of trophoblast cells (191, 209, 539). HPV detection on spermatozoa was predictive of negative pregnancy outcome, whereas in infertile couples, HPV vaccination of the male partners with HPV in their semen significantly improved the pregnancy and live birth rate compared with the nonvaccinated group (212). Further studies on the consequences of sperm washing and removal of HPV-bound spermatozoa on assisted reproduction outcome in large cohorts of infertile couples with HPV+ men are warranted, to substantiate the negative impact of HPV+ spermatozoa on reproduction.
The vertical transmission of teratogenic ZIKV by spermatozoa is a possibility that merits investigation since infectious virus and viral proteins were detected in motile spermatozoa (326, 427). In a mouse model, experimental infection of ova with teratogenic CMV used to mimic sperm-mediated infection led to fetal anomalies upon transfer to surrogate mothers (44). HHV-6 specifically integrates into telomere regions of human chromosomes, establishing life-long latency. HHV-6 is endogenous in ~1% of the human population, and vertical transmission of chromosomally integrated HHV-6 from either the father’s spermatozoa or from the mother has been demonstrated (26, 130, 273, 479, 672, 725), leading to virus integration in every cell of the offspring (273, 725). For at least one subset of individuals with chromosomally integrated HHV-6, endogenization was estimated to have occurred 14,000–35,000 yr ago (769). Whether HHV-6 invasion of the male gametes continues to happen in present times, and the impact of such integration, is currently unknown.
Significantly, medically assisted procedures such as intra-cytoplasmic sperm injection (ICSI) and in vitro fertilization (IVF), which overcome physiological barriers, can be at increased risk of virus transfer from infected sperm/testicular germ cells to the oocyte if infected gametes remain. The zona pellucida, which surrounds the oocyte, and the oocyte membrane, although not impermeable, constitute natural barriers against viruses bound to the surface of spermatozoa that are bypassed during ICSI. The whole sperm/testicular germ cell, including its cell membrane, is injected into the oocyte cytoplasm, instead of the physiological requirement for zona pellucida crossing and cell membrane fusion between the male and female gametes. Also, the motility and fertilization ability of spermatozoa can be impaired by the infection (see sect. VIB1), which represent another natural protection against viral contamination during fertilization that is broken by IVF and ICSI.
Finally, male gametes have the ability to take up foreign DNA and might therefore pass on viral gene fragments to the offspring. This property led to the development of many studies aiming to make transgenic animals using “transgenic spermatozoa” as vectors. However, the results have proven disappointing overall, essentially since most foreign DNA remains extra-chromosomal, leading to mosaicism in the offspring (89).
C. Viruses as Invaders: Persistence of Emerging and Contemporary Viruses in the MGT
1. Zika virus
During the major 2015–2016 ZIKV outbreak in the Americas, high viral loads and prolonged shedding of ZIKV RNA (up to 1 yr) and infectious particles in semen (up to 69 days) from infected men, despite viral clearance in the blood, supported the existence of ZIKV reservoirs within the MGT (43, 169, 452). Worryingly, seminal shedding, persistence, and sexual transmission (the latter up to 41 days post symptoms) occurred even in asymptomatic patients (169).
The deleterious impact of the infection reported on sperm parameters, along with the association of ZIKV with ejaculated spermatozoa and exfoliated testicular germ cells, suggested infection of the testis and epididymis (307, 326, 427, 441). In human testis tissue infected ex vivo, ZIKV replicated in a wide range of testicular somatic cells: resident macrophages, peritubular cells and to a lesser extent Leydig cells and Sertoli cells, as well as immature and mature germ cells (441). Infection of the testis tissue did not trigger cytopathic effect or IFN upregulation (needed to sustain an efficient antiviral defense), and induced only minimal pro-inflammatory cytokine stimulation, similar to the persistently infected cerebrospinal fluid of Indian macaques (5). IFNs play a key role in inhibiting ZIKV replication, and its production in humans is counteracted by the virus. However, the lack of IFN upregulation following viral exposure could be a feature of the human testis that is not specific to ZIKV (692 and unpublished data). Altogether, the survival of infected cells and weak innate immune response may represent favorable conditions for ZIKV persistence in the human testis (441). ZIKV infection persisted for up to 6 wk in a human Sertoli cell line, leading to a relatively muted innate response (665). In immuno-competent mice, ZIKV preferentially infected testicular germ cells over somatic cells, and persisted in the testis for up to 60 days (581). Rodent meiotic and post-meiotic germ cells are known to have impaired IFN signaling (136, 602), increasing their susceptibility to ZIKV infection compared with other testicular cell types (581). ZIKV replication did not affect mouse germ cell survival or proliferation (581). In mice with deficient IFN signaling (e.g., IFNAR−/−), a broader range of testicular cells was targeted by ZIKV as with human testis (441), and a high level of ZIKV was detected up to 42 days post-infection in both testis and epididymis (96, 247, 410, 696, 702). In IFNAR−/− mouse epididymis, the virus was associated with luminal sperm, macrophages and epithelial cells (696). In NHP, multiple anatomic sanctuaries of ZIKV were identified, including the central nervous system, lymph nodes and, at relatively low levels with inter-individual variations, testis, prostate, and seminal vesicles (5, 287, 524, 657).
Vasectomy in mice and men significantly lowers ZIKV viremia in semen, which indicates that testis/epididymis are important viral sources (157, 452). However, vasectomy does not abrogate sexual transmission despite viral clearance in blood (30, 157). It is noteworthy that the longest duration of infectious ZIKV in semen (69 days) was observed in a vasectomized man, and that ZIKV RNA persisted in semen from vasectomized individuals up to 281 days post symptoms (30, 195, 307, 452). While divergent results were obtained regarding prostate and seminal vesicle infection in either immune-deficient mice or NHP (657), it is evident that MGT reservoirs other than testis and epididymis exist in ZIKV-infected men. How and which MGT organs other than testis can shelter ZIKV from immune clearance warrants investigation pertaining to the immune status of these organs and to the nature of the reservoir cell types.
2. Other arboviruses
Case reports of excretion in semen after systemic clearance of yellow fever virus (YFV) (40), dengue virus (DENV) (377), and Chikungunya virus (CHIKV) RNAs (38) for up to 19 days, 37 days, and 30 days post-symptom onset, respectively, suggest that several arboviruses may persist in the MGT (TABLE 7). Furthermore, WNV was recently suspected to be sexually transmitted 30 days post-symptoms onset (348), along with sexual transmission of DENV between two men (391). This concern is further substantiated by the persistence of arboviruses (e.g., Japanese encephalitis virus, JEV) in animal semen and MGT organs (see sect. VIII). There is currently no data on human MGT organs infection for these viruses. In mouse models, DENV did not productively infect testicular cells (247, 410, 581, 627). In contrast, YFV productively infected a human Sertoli cell line (634). The duration of viral excretion following peripheral clearance and the demonstration of infectious virus in semen are still to be explored in human cohorts to determine whether the MGT is a reservoir for these arboviruses.
Prevalence of Virus Shedding in Semenb | ||||||
---|---|---|---|---|---|---|
Persistent semen infection despite undetectable viremia | ||||||
Virus | Early stage of infection | Chronic stage of infection | Max. duration reported/% of persistent shedders | Median time until RNA clearance | Shedding Pattern (%Continuous) | Sexual Transmission Reported |
HSV-1, -2 | NA | 0–10%c,e (225, 227, 229, 394, 482, 523) | 50% (482) | + | ||
VZV | NA | NA | NA | |||
EBV | NA | 4–56%c,e (225, 227, 229, 394, 482, 523) | 66% (482) | + | ||
CMV | NA | 4–70%c,e (203, 225, 227, 229, 230, 394, 482, 523, 628) | 81% (482) | + | ||
HHV-6 | NA | 1–7%c,e (225, 227, 229, 394, 482, 523) | 46% (482) | + | ||
HHV-7 | NA | 6–15%c,e (225, 227, 229, 394, 482, 523) | 57% (482) | |||
KSHV | NA | NA | 50% (482) | + | ||
HIV-1 | 61–100%b (104, 227, 228, 482) | 81–100%b (313, 523, 628) | d1789 (13.4%) (123, 223, 229, 298, 533) | 9 days (313) | Chronic: 56–61% (115, 265); ART: 50–75% (80, 272, 629) | + |
HBV | NA | 68%c,d (180) | d120 (2 cases) (177, 310) | NA | NA | + |
HCV | 29–39% (71, 700) | 32–46%c,d (71, 73, 386, 532, 700) | 0–28% (73, 532) | + | ||
EBOV | 73–100%a,b (133, 583, 591, 640, 704) | d1178 (0.4%) (646) | 4 mo (640) | 100% (41, 640, 704) | + | |
ZIKV | 50–68%a,b (43, 307, 381, 452) | d370 (10%) (43) | 25–83 days (43, 307, 452, 536) | 100% (30, 129, 195, 207, 381, 493, 506) | + | |
HPV | 11.4%c,f (409) | d730 (15%) (85) | 15 mo (85) | NA | + | |
AAV | 0–5%c (210) | + | ||||
SV40 | 33–45%c (431, 432) | |||||
JCPyV | 0–21%c (112, 432, 472) | |||||
BKPyV | 0–95%c (112, 431, 432, 472) | |||||
GBV-C/HGV | 15%c (617) | |||||
TTV | 59%c (439) | |||||
MARV (case report) | + (433) | + | ||||
HTLV-1 (case report) | + (318) | + | ||||
WNV (case report) | + (348) | |||||
YFV (case report) | NA | d21 (40) | NA | |||
DENV (case report) | NA | d37 (377) | NA | + (391) | ||
CHIKV (case report) | NA | d30 (38) | NA | |||
LFV (case report) | NA | d103 (566) | NA | |||
RVFV (case report) | NA | d117 (277) | NA | |||
ANDV (case report) | NA | d278 (372) | NA | |||
NiV (case report) | NA | d26 (31) | NA | |||
MuV (case report) | NA | d40 (319) | NA | |||
SFV (case report) | + (61) |
3. Ebola and Marburg viruses
The ability of filoviruses to persist for extended periods in the MGT was reported in the late 1960s for the first human cases of EBOV and MARV infection (167, 434), but the issue only came to light properly during the major 2014–2016 West Africa outbreak of EBOV (28,000 cases and 17,000 survivors). EBOV RNA was found in the semen of survivors for as much as a remarkable 1,178 days (646), and proven sexual transmission from male survivors to female partners up to 470 days after disease offset (609). Alarmingly, persistent infection of the MGT by EBOV is asymptomatic. Sexual transmission cases from survivors distant from the epidemic contributed to its resurgence through the initiation of new transmission chains (609). Fortunately, as for ZIKV, filovirus persistence in MGT is not lifelong and steadily declines over time until the virus eventually disappears, in most cases within months (WHO currently recommends a 9-mo abstinence in male EBOV survivors), but years for a subset of individuals (609). In a study on 267 Ebola survivors, a positive association between the presence of uveitis, an inflammation of the eye, and detection of viral RNA in semen samples was found. Since uveitis was associated with higher viral load during the acute stage of the disease, this may suggest that high viral load at the onset of the infection is associated with seeding of the MGT and in turn prolonged excretion in semen (646).
Higher seminal viremia during recovery, compared with blood viremia at peak illness, suggested persistent active replication of EBOV within the MGT. This was further demonstrated through the detection of positive sense RNA (i.e., nongenomic, replicative form) in semen (41, 730). The low rate of change or absence of change in EBOV genome in longitudinal semen samples (41, 640, 730) suggests reduced immune pressure compatible with infection of immune-privileged sites (note that similar low rates were found in the eye, another immune-privileged site of persistence) (730).
The nature of the reservoir(s) in the MGT which seed EBOV into semen is currently unknown. EBOV antigens were present in the testicular interstitium (including endothelial cells) and seminiferous tubules (suggestive of Sertoli and germ cell infection) from a fatal human case, with no histological lesions or inflammation (430). In acutely infected rhesus and cynomolgus macaques, EBOV was localized in the vascular structures, endothelial cells, and mesenchymal cells within the interstitial tissue of the testis, epididymis, prostate, and seminal vesicles, with no changes in organ morphology (541, 766). In a subset of animals with delayed time of death, EBOV infected the seminiferous tubules, with a staining suggestive of Sertoli cell infection (766). In one NHP survivor, EBOV was associated with macrophages in the lumen of the inflammatory epididymis (766). EBOV was not detected in the testis, but animals were sexually immature (spermatogenesis had not started), which may have impacted EBOV tropism in the MGT. EBOV persisted in the testis of humanized mice that survived the infection (57), and infected the testicular interstitium of experimentally infected guinea pigs (113). In addition, EBOV infected clusters of epithelial cells as well as stromal macrophages and fibroblasts in the penis and foreskin of acutely infected guinea pigs (116). In semen, EBOV RNA and infectious particles were rescued from seminal plasma in several studies (609). To the best of our knowledge, there has been no attempt to identify EBOV-infected cell types in semen. This information might help to shed light on the origin of persistent EBOV in semen.
MARV was the first filovirus reported to be sexually transmitted. In 1967, a man with MARV antigens in semen cells contaminated his wife during sexual intercourse 2 mo after recovering (434). Infectious MARV has been rescued from semen up to 84 days post-recovery (434, 509, 643). In contrast to EBOV, MARV induced testicular inflammation in infected individuals, as evidenced at autopsy and through reports of orchitis, testicular pain, and swelling up to 52 days post-infection (609). In acutely infected NHP, MARV was localized in the testis interstitium and endothelial cells, as well as in the epididymal smooth muscle cells and prostate stroma (117). In the testis of NHPs that survived MARV, inflammatory infiltrates were localized specifically in infected areas within the interstitium and seminiferous tubules and were associated with BTB disruption and focal germ cell depletion (111). Overall sperm production was not affected. Similar to observations in mice infected with ZIKV (441, 696), MARV arising from testis blood vessels first infected interstitial Leydig cells and peritubular cells, and in a second phase persistently infected seminiferous tubules. Sertoli cells supported active viral replication and were the main viral reservoir within the testis, with only low numbers of germ cells and macrophages infected. Focal testicular immune infiltrates were composed of T and B lymphocytes, macrophages and neutrophils. T cells were essentially CD4+ (T helper-1, -2, -17 and Tregs) with a few CD8+ cells. Interestingly, the presence of Tregs was specifically associated with persistence, along with a high TGF-β concentration and cytotoxic T-lymphocyte-associated protein 4 (CTLA4) expression, suggesting localized immunosuppression, which may delay clearance of MARV in the testis (111). Infectious viral particles released by productively infected Sertoli cells in the seminiferous tubule lumen could be a source of MARV in semen. MARV tropism for human Sertoli cells now needs to be established.
4. Human immunodeficiency virus
The persistence of HIV-1 RNA and DNA was reported in the semen of a subset of HIV-infected men receiving efficient antiretroviral therapy (ART), despite undetectable blood viral loads for several years (298). In experimentally infected macaques with controlled blood viremia below 100 copies/ml during ART, SIV persisted in semen from a subset of animals after 4 mo of treatment (440). HIV-1 and SIV infect a broad range of MGT organs and ducts (testis, efferent ducts, epididymis, vas deferens, prostate, seminal vesicles, urethra, foreskin and glans), as shown at autopsy from men deceased during AIDS (511, 557) or in the chronic asymptomatic stage (200, 322, 486), in experimentally infected NHPs during the acute (474, 693), chronic (299, 440, 625, 693), or AIDS stage (465), and using ex vivo-infected human MGT tissue (141, 201, 590, 694). Throughout the MGT, infected cells are composed of tissue-resident macrophages and/or CD4+T lymphocytes (298). In the testis, HIV/SIV was associated with isolated germ cells in a few studies (486, 511, 625, 693), with no spermatogenesis disruption. In SIV-infected macaques, the urethra was the most highly infected MGT organ, and the testis was the least infected, which could reflect a much higher number of immune target cells in the urethra than the testis (440). While the local sources of HIV in semen remain elusive (298), a recent study by our laboratory in the SIV-infected macaque model demonstrated that multiple MGT organs shed virus particles and infected cells in semen, as identified through phylogenetic analysis (299). Surprisingly, viral excretion from distinct organs was independent from inflammation and organ infection level, and varied among individuals (299).
Only a few studies attempted to decipher the organ reservoirs for HIV/SIV in the MGT of treated individuals. Our laboratory first revealed in SIV-infected macaques under ART for 4 mo the persistence of SIV RNA+ macrophages in the urethra, whereas productively infected cells were no longer detectable in other MGT organs (440). In agreement, in HIV-infected individuals under suppressive ART for years, urethral macrophages localized in the stroma and epithelium harbored integrated replication-competent virus. Virus production could be reactivated by bacterial lipopolysaccharide (LPS), albeit at a low level. Virions were contained within intracellular compartments called VCCs (virus containing compartments) (200). VCCs were previously evidenced in blood-derived macrophages, allowing viruses to retain their infectivity for extended periods of time (582). Infected urethral macrophages had an intermediate M1-M2 polarized macrophage phenotype, which was enriched in HIV-infected men. The authors postulated that the high urethral concentrations of IFN-γ could inhibit the killing ability of the cytotoxic T lymphocytes found associated with the infected macrophages (200). Although the urethra represents a reservoir for HIV, persistent HIV excretion in semen appears to involve different MGT; thus, in 10 treated HIV-infected men with persistent HIV in the semen, no virus was detected in the pre-ejaculatory fluid produced in the urethra (547). Multiple MGT organs could act as reservoirs. Indeed, we showed in macaques that received 4 mo of ART that SIV DNA levels in testis, epididymis, vas deferens, seminal vesicles, and prostate were similar to those of untreated animals (440). Similarly, HIV DNA+ cells persisted in the testis from treated men (322), suggesting the persistence of latently infected cells and the potential for reactivation. The organs excreting the virus in semen probably vary between individuals, as we demonstrated in macaques (299). Specific tissue micro-environments may play determining roles in the reactivation of latently infected cells and in infection susceptibility of long-lived/drug-resistant cells, such as the highly plastic tissue macrophages.
5. Human papillomaviruses
In men, most genital HPV infections are subclinical and clear within 1–2 yr post-infection (236). HPV persistence in the penis and urethra has been investigated (85, 375), but results were complicated by the fact that the sexual partners of the male participants were also HPV positive, and therefore persistence could not be discriminated from new infection. In one study on elderly men who no longer had sexual intercourse, the same HPV subtype was estimated to persist in the skin of the glans penis for an average of 8 yr in the absence of any new source of infection (359). Higher loads of HPV in penile samples conferred a greater probability of viral persistence of up to 12 mo but not more, suggesting other factors as yet unidentified for long-term persistence in the MGT (398).
6. Herpes simplex virus type 2
An early study suggested that the MGT serves as a silent reservoir for HSV-2, based on the high incidence of actively replicating virus in the vas deferens and prostate from men of different age groups (up to 70+), in the absence of any history of genital herpes together with low frequency of detection in the urethra (93). HSV-2 infects genital epithelial cells during primary infection and spreads to the sensory neurons, where it establishes latency (363). HSV-2 reactivation is accompanied by an influx of both CD4+ and CD8+ T cells and is normally contained within days (608). High numbers of Tregs in HSV-2 reactivation areas were correlated with increased viral replication (466). It would therefore be interesting to investigate whether actively replicating HSV-2 in MGT organs is associated with elevated Treg numbers which impairs immune control, as postulated for MARV infection in the testis (111).
VII. SEMEN, A TROJAN HORSE FOR THE SEXUAL TRANSMISSION OF VIRUSES
Over 30 viruses were found in human semen, of which 15 are associated with sexual transmission (TABLE 7). On top of well-known sexually transmitted viruses such as genital herpes viruses and blood-borne HIV, new additions include vector-borne ZIKV and blood-borne EBOV, which were shown to persist in semen for an extended period after blood clearance. Out of 16 case reports of various viruses in semen, 7 were reported as persistent (TABLE 7).
A. Prevalence and Pattern of Virus Shedding in Semen
The prevalence, level, and duration of shedding in semen vary widely for different viruses and significantly impact the dynamics of an epidemic. During early infection associated with high viremia, there is frequent seminal shedding of HIV, HBV, HCV, ZIKV, and EBOV, ranging from 29 to 100% (TABLE 7). Similar shedding frequencies are observed in the chronic stages of HIV, HBV, and HCV despite the lower viremia (TABLE 7). HIV load in semen is usually lower than in blood, but patients with an exceptionally high semen viral load were reported (533, 671, 741). Factors associated with HIV shedding prevalence in semen are blood viral load (104, 176, 267, 313, 329), STIs (329, 548), seminal CMV or EBV (226, 227, 229, 394, 482, 523, 628), and seminal cytokine levels (394, 395, 523, 548), the two latter reflecting elevated inflammation, which enhances HIV replication.
In most studies, HCV load in semen is positively correlated to viremia (71, 73, 700), and a higher prevalence of HCV shedding in semen from HIV-infected patients is controverted (71, 73). However, Turner et al. (700) only observed a correlation between blood and semen HCV loads during the acute stage, suggesting that other factors such as concomitant shedding of human herpesviruses (HHVs) could affect HCV shedding.
Persistent seminal shedding designates the long-lasting excretion in semen of viruses that have been controlled or cleared in the periphery by antiviral treatments (e.g., HIV, HBV) or by the immune system (e.g., ZIKV, EBOV). It implies an undetectable viremia and strongly suggests that seminal viruses and infected cells originate from the MGT (see sect. VIC and below). In HIV-infected patients, the virus level in semen is usually severely decreased in the first weeks following the start of antiretroviral treatment (313). However, HIV RNA persists in semen for months or years post-treatment initiation in ~13.4% (range 1.8–48%) of patients, despite undetectable viremia (223, 229, 298, 533). In contrast, there are no cases of HCV persistence in semen following treatment and only two controverted cases of HBV persistence in semen in the absence of viral DNA in serum (177, 310, 351).
Viral RNA was detected in semen from ZIKV- and EBOV-infected patients until day 370 and day 1,178, respectively, after symptom offset (TABLE 7). In ~50% of EBOV patients, the virus disappeared from the semen between 3 and 6 mo post-infection (133, 640, 649), and in all studies, the EBOV titer in semen steadily decreased over time (41, 640, 649, 704). Nevertheless, 8% of a cohort of 137 EBOV patients still shed virus in semen 2 yr after the disease (187), with older age associated with seminal shedding (187, 646, 649). For ZIKV, the mean clearance time in semen was 25–83 days, as opposed to 5–15 days in blood (43, 189, 307, 452, 536). Age, absence of ejaculation, conjunctivitis, and joint pain were factors associated with prolonged virus shedding (452). Why older age would influence seminal shedding duration is unknown. Reduced immune functions and endocrine or anatomic (e.g., prostate hyperplasia) modifications might be involved.
HPV is responsible for genital infections that usually clear within 2 yr. In line with this, the median duration of HPV detection in semen samples was 15.3 mo (85). HHVs establish lifelong infection and are characterized by a high worldwide prevalence, ranging between 20 and 100% for all HHVs except KSHV. All HHVs were reported as being present in semen at least once (TABLE 7). The reported prevalence of 37.3% for all HHVs in the semen of healthy patients (482) increased to 59–92% in HIV-infected patients (225, 227, 394, 482). Considerable variations in semen prevalence exist between the herpes viruses, but also across studies, depending on geographic location, detection method, and type of cohort, mostly HIV-infected and uninfected patients or fertility-clinic attendants (336). Overall, the semen prevalence of HHVs in HIV-negative men is under 15%, while HIV infection most dramatically increases CMV and EBV seminal shedding (TABLE 7) (203, 224–227, 229, 230, 394, 482, 523, 628).
Longitudinal studies reported a mix of intermittent and continuous shedding patterns for HIV and HHV during the early and chronic stages of the disease (TABLE 7). The link between virus shedding and other STIs, herpes/HIV co-shedding, and seminal cytokine levels suggests that local factors related to transient infection/re-activation of virus sources could be involved in this intermittent process. However, during undetectable blood viremia in ART-treated patients, the HIV shedding pattern was intermittent, with intervals as little as 1 h between positive and negative samples (181). Such short intervals suggest that intermittence may also reflect the natural variation of semen composition between ejaculates (422). Unlike HIV and HHVs, the shedding pattern during the persistent stage of EBOV and ZIKV is generally continuous (TABLE 7), with occasional intermittent detection most likely due to the assay sensitivity threshold (187). This suggests that the main factor driving the shedding pattern is poor efficiency of the local immune system to control ZIKV replication in the MGT. This is supported by the detection of pro-inflammatory cytokines together with infectious virus in seminal fluid long after viral clearance from serum (426, 519).
B. From Where Do Virions and Infected Cells in Semen Originate?
Systemic viruses that contaminate semen during the acute infection/high blood viremia stage most likely arise from the passive diffusion of viral particles and infected cells from the blood. In favor of this hypothesis are concomitant high blood and semen viremia in acute HIV infection, together with similarities in nucleic acid sequences of the blood and seminal HIV and SIV strains (298). In contrast, several studies reported an absence of association between semen and blood viral loads during the chronic stage of HIV infection (298). Phylogenetic studies comparing viral populations in blood and semen showed differences in over 50% of HIV-infected patients and SIV-infected macaques (298), suggesting a local origin within the MGT. This was recently demonstrated by Houzet et al. (299) who showed, based on the phylogenetic comparison of viral strains in blood and semen fluids and cells with that in male genital organs from macaques chronically infected with SIV, that seminal virus and infected cells originated from various genital organs such as the seminal vesicles, vas deferens, and epididymis. With regard to genital HPV, the excretion of virus in semen was associated with penis epithelium infection, probably reflecting the exfoliation of infected epithelial cells (407).
The absence of viremia during the persistent seminal shedding of HIV, ZIKV, and EBOV strongly suggests a local origin at this stage per se. The presence of infected testicular germ cells and spermatozoa in the semen of ZIKV-infected patients points to the testis and/or epididymis as the sources of infected cells (326, 427, 441). Nevertheless, the persistence of ZIKV RNA and infectious viral particles in the semen of vasectomized patients indicates that distal MGT organs (e.g., prostate, seminal vesicles, and urethra) constitute additional viral sources (30, 195, 307, 452).
C. Importance of Semen Contamination for Virus Dissemination
Fifteen of the over 30 viruses detected in semen are sexually transmitted (TABLE 7), including HIV, HSV-2, HPV, and HBV, which are responsible for four major sexually transmitted diseases. Sexual transmission is also an important mode of transmission for HTLV-1 in endemic areas (526). Semen is a key vector of transmission for HIV-1, HSV-2, and HTLV-1, with more efficient men-to-women transmission than women-to-men (526, 744). While the sexual transmission of HCV has long been considered negligible, seminal HCV is now thought to play a significant role in HCV spreading in MSM (700). For viruses like EBOV and ZIKV, sexual transmission is a minor mode in endemic countries compared with other modes (i.e., other body fluids, fomites, and mosquitoes). However, recent work reported transmission via semen as the source of multiple flare-ups during the last EBOV outbreak (62, 144), and mathematical models integrating the sexual transmission mode demonstrate a significant impact on epidemic dynamics (2, 406, 714). Similarly, computational models show that sexual transmission of ZIKV might be underestimated (14) and could impact on the size of the epidemic and length of the outbreak (204). Recently, the analysis of risk factors in 336 household contacts of ZIKV patients showed an increased risk for sexual partners in an endemic context (587). For both EBOV and ZIKV, long-lasting virus in semen represented the main issue regarding sexual transmission. Vertical transmission of viruses such as ZIKV through semen, as demonstrated in mice (156), is another important issue to consider.
Accumulating case reports demonstrate that a number of (re)emerging viruses can infect semen (e.g., NiV, YFV, DENV, CHIK), some for extended durations (e.g., 278 days for ANDV, 117 days for RVFV, 103 days for LFV) (TABLE 7). This raises great concerns since some of these viruses are lethal and detected in the semen of survivors (e.g., NiV, LFV). Sexual transmission was recently suspected for WNV 30 days after symptom onset, and sexual transmission of DENV between two men was reported (TABLE 7). Whether sexual transmission of the emerging viruses that contaminate human semen can occur warrants study.
D. Semen Is More than a Passive Vector for Viruses
Viruses can be present in semen as cell-free virions, in infected cells (e.g., leukocytes infected with HIV and spermatozoa infected with HBV), and/or attached to cell surfaces (e.g., HPV and HIV bound to spermatozoa). Both cell-free and cell-associated virus transmission occurs through the anogenital mono- or pluri-stratified mucosa of the recipient (20, 245), involving a number of distinct mechanisms for crossing the mucosal epithelial barrier (FIGURE 11). Over the last two decades, the accumulated in vitro data have shown that seminal plasma exerts a complex mix of inhibitory and enhancing effects on viral infection, depending on the target cells and pathogens. It is important to note that some of these seminal effects may be transient, since seminal fluid is a dynamic fluid, the composition of which changes over time post-ejaculation due to many enzymatic reactions.

FIGURE 11.Mechanisms for sexual transmission of viruses through the recipient mucosal epithelia (A–E) and modulation by seminal plasma components. Viruses present in genital secretions (semen, cervico-vaginal fluid) as virions and within infected cells can be transmitted through the monostratified (e.g., endo-cervix, colon, penile urethra) or pluristratified (e.g., vagina, ecto-cervix, foreskin, glans) mucosa of the recipient through a range of mechanisms. A: for sexually transmitted viruses with a tropism for epithelial cells such as herpes simplex virus (HSV) or human papillomavirus (HPV), the epithelium that covers mucosal tissues is productively infected, providing a direct entry way. Viruses that do not infect epithelial cells, such as human immunodeficiency virus (HIV) or human T-lymphotropic virus, may reach their subepithelial target cells through either breaches in the mucosa resulting from micro-abrasions of the epithelial barrier triggered by sexual intercourse or by other sexually transmitted infections favoring mucosa inflammation or abrasions (B), transmigration of infected cells between the epithelial barrier (C); capture of viral particles by dendritic cells sampling pathogens on the apical side of the epithelium barrier, which then migrate back into the subepithelium tissue (D); or transcytosis of viral particles across the epithelial cells leading to their release on the basal side (E). F: transmission of free viral particles captured by DCs to permissive submucosal cells. For detailed reviews of these mechanisms, see References 21, 626. Multiple semen components (showed in grey clouds of semen in the schema) are thought to enhance (green arrows) or inhibit (red bars) viral transmission (see sect. VIID for more details on their mechanisms of action). ROS, reactive oxygen species; TGF-b, transforming growth factor-β; PGE2, prostaglandin E2; HSPG, heparan sulfate proteoglycans; ZIKV, Zika virus; DENV, Dengue virus; WNV, West-Nile virus; CMV, cytomegalovirus; EBOV, Ebola virus.
1. Inhibitory effects of interactions between semen components and viral particles or their target cells
a) semen exosomes.
Also called prostasomes, exosomes are membranous nanovesicles produced within MGT organs, the content (proteins, mRNAs, and miRNAs) of which is involved notably in sperm maturation and fertilization. Seminal plasma and isolated semen exosomes blocked the binding of ZIKV, DENV, WNV (487), and CMV (393) to their target cells and inhibited HIV transcription and cell-to-cell spread (166).
b) cationic polypeptides.
In high concentrations in semen, several cationic peptides demonstrated antiviral activity in vitro. Similar to seminal plasma, a natural fragment of semenogelin I transiently inhibited HIV infection of the target cells (429). Seminal vesicle-derived gp17 glycoprotein specifically competed with HIV for binding to its cell entry receptor CD4 (33), while a cysteine proteinase inhibitor, cystatin C, was postulated to counteract viral proteases necessary for viral protein processing (164, 709).
c) reactive oxygen species.
Reactive oxygen species (ROS), mostly produced by neutrophils and macrophages in semen, modified lipid rafts in the membranes of enveloped viruses including HIV, in turn inhibiting fusion with the cell and viral entry (290, 361).
d) clusterin and mucin 6.
The glycan motifs of clusterin and mucin 6 bind to DC-SIGN, a PRR of the C-type lectin family expressed by genital mucosa immature DCs, which recognizes many viruses (HIV, HSV, HCV, DENV, ZIKV, EBOV, CMV) (276, 367). These glycoproteins competed with HIV for binding to DC, in turn inhibiting trans-infection of CD4+ T cells, similar to whole seminal plasma (597, 659).
e) rantes/ccl5.
Elevated concentrations of the chemokine RANTES, a ligand of the HIV receptor CCR5, were found in semen from HIV+ men and diminished HIV infection of semen-exposed CD4+ T cells by decreasing their CCR5 expression (84).
2. Enhancing effects of semen components/properties on virus transmission
a) amyloid-forming fibrils.
Over 10 yr ago, Munch et al. (488) identified seminal amyloid-forming peptides that boost HIV infectivity in vitro, similar to whole seminal plasma. These peptides comprise fragments of the prostatic acid phosphatase (with a dominant peptide called SEVI for semen-derived enhancer of viral infection) (488) and seminal vesicle-derived semenogelins (SEM1 and SEM2, resulting from cleavage by the prostate specific antigen) (577, 578). The positive charges of the peptides interact with the negatively charged surfaces of cells and virions to form an electrostatic bridge that promotes viral attachment and fusion (29, 578). Seminal plasma, SEVI, and semenogelin amyloids enhanced in vitro infection of isolated cells by HIV-1 (355, 488), HIV-2 (355), SIV (775), HSV (690), CMV (673), and EBOV (42) but not that of ZIKV, WNV, and DENV, which were inhibited by seminal plasma (487). SEVI also increased the internalization of EBOV particles by macropinocytosis, the canonical EBOV entry pathway, and stabilized EBOV viability and infectivity (42). However, seminal plasma or SEVI enhancement of HIV infection in ano-genital explants was inconsistent (15, 315, 368), and in vivo experiments using humanized mice (120, 148) or NHP models (489) failed to demonstrate any enhancing effect of semen on HIV/SIV transmission. This discrepancy could result from the complex interactions between seminal and mucosal factors counteracting amyloid fiber enhancement. It is also possible that the higher viral doses used in vivo/ex vivo compared with in vitro may bypass the semen-enhancing effect, which is restricted to low viral doses (488).
b) cytokines.
The role of seminal cytokines in viral transmission is ambiguous. Semen has very high concentrations of TGF-β (mostly under a latent form activated by vaginal pH and enzymes) and prostaglandin (PG) E2, which both inhibit leukocyte activation and induce naive CD4+ T cell differentiation into Tregs (580). Semen immunosuppressive properties, believed to allow maternal tolerance against paternal-derived antigens, might facilitate virus spreading. In addition, TGF-β increased CD169 expression on mature DC, which facilitated HIV capture by DCs and may enhance its transmission to permissive cells (599).
Conversely, semen deposition onto the female genital mucosa triggers a strong post-coital inflammatory response with massive influx of neutrophils, and to a lesser extent lymphocytes and macrophages, which may eliminate pathogens (580). PGE2 stimulated neutrophil chemotactic CXCL8 in cervix tissue (142), while seminal plasma, PGE2, and TGF-β upregulated COX2 in female genital cells (315, 328, 623), which may further amplify the inflammation by catalyzing prostaglandin synthesis. Compared with blood, semen contained elevated concentrations of pro-inflammatory cytokines [e.g., MCP-1, IL-8, CXCL1, MIP1-β, IL-6, IL-1α, granulocyte-macrophage colony-stimulating factor (GM-CSF)] and adaptive cytokines (IL-7 and IL-15) (522, 707). In cervix and endometrial tissue cultures and in vaginal, cervical, and endometrial isolated cells, seminal plasma upregulated a range of genes and proteins linked to cytokine signaling, inflammation, antigen presentation, leukocyte migration, and cellular immune responses, a modulation in part mediated by TGF-β (99, 315, 622). Unfortunately, semen cytokine-mediated inflammation and immune activation attract HIV immune target cells and stimulate replication of the virus in female genital tissue (50, 315).
Semen upregulated CCL20 chemokine in cells derived from endocervical and vaginal epithelium, possibly through the action of seminal lactoferrin (403), which in turn recruited CCR6+ Langerhans cells in vitro (50). In NHPs, CCL20 concentrations were positively correlated to recruitment beneath the vaginal epithelium of CCR6+ plasmacytoid DCs and macrophages, which in turn attracted T helper 17, a preferential target cell type for HIV/SIV (662).
c) fibronectin.
Fibronectin, an abundant extracellular matrix glycoprotein that stimulates sperm capacitation and semen coagulation, coats HIV virions (69). Fibronectin recognition of B1 integrin and HSPGs expressed by epithelial cells might help HIV attachment to the female genital epithelium (419, 576). Multimeric fibronectin enhanced the HIV infection of T lymphocytes through increased cell attachment and protected virions from degradation (254, 678).
d) complement molecules.
The opsonization of virions coated by complement molecule present in semen by epithelial cells, DCs, and macrophages bearing the CR3 complement receptor was postulated as an enhancing mechanism for HIV transmission (66).
e) semen pH.
The alkaline pH and high buffering capacity of seminal fluid protected HIV particles from the acidic environment of the vaginal mucosa, in turn stimulating infection (68).
VIII. WHAT LESSONS CAN WE LEARN FROM ANIMALS?
Of the 87 new human pathogens catalogued since 1980, ~80% are zoonotic (498). Considering the potentially high impact of sexual transmission of viruses in upcoming human pandemics, and the urgency to fill gaps in this domain revealed by the recent ZIKV and EBOV epidemics, it is essential that we learn from diverse research fields. This includes research into economically significant farm animals (e.g., bulls, pigs, sheep, and horses). Since artificial insemination is used extensively worldwide in these animals, viruses that infect and persist in the MGT represent a threat for the disease spreading and breeding and have been sought intensively. The main findings on viruses in the MGT of farm animal are summarized below and in TABLE 8 to learn from the mechanisms of seminal viral shedding and persistence evidenced and prompt research into viruses that might infect the human MGT.
Host | Viral Family | Virus | MGT Organs and Cells Infected | Seminal Excretion (S) and Persistence (S+) | Venereal Transmission (V), Reproduction Failure (R), Teratogen (T), Abortion (A), Embryo Death (E) |
---|---|---|---|---|---|
Swine | Anelloviridae | TTV | T (574) | S+ (347, 574) | |
Asfarviridae | ASFV* | S (680) | |||
Circoviridae | PCV2a | T, E, SV, P, BG (250, 416, 574) | S (380, 416) | R | |
Flaviviridae | CSFV* | T+ (germ cells), E+, VD+ (107) | S+ (108) | E | |
Flaviviridae | JEV*†a | T (268) | S (400) | ||
Flaviviridae | BVDV | T+b, P+b (Mϕ), SV+b (Mϕ) (679) | S+b (679) | ||
Flaviviridae | APPV | S (614) | |||
Parvoviridae | PPV | E (248) | S (354) | V?; R | |
Parvoviridae | PPV4 | S (126, 216) | |||
Herpesviridae | PRV | T, E, foreskin (274, 467) | S (453) | ||
Paramyxoviridae | PoRV | T+, E+ (575, 650) | S+ (575) | R | |
Arteriviridae | PRRSV | T (germ cells, macrophages), E, VD, SV, PR (Mϕ) (109, 632) | S+ (109) | V; R | |
Coronaviridae | PEDV | (ND) | S (199) | V? | |
Picornaviridae | PEV | S (544) | |||
Picornaviridae | PTV | S (544) | |||
Picornaviridae | FMDV | S (451) | |||
Picornaviridae | SVDVa | S (451) | |||
Hepeviridae | HEV†a | S (388) | |||
Bull | Flaviviridae | BVDV | T+ (Sertoli, germ cells, epithelial cells), P+, E+, SV+, U+ (epithelial cells, fibrocytes) (63, 238, 358, 504, 716) | S+c (63, 358) | R; E; T; A |
Flaviviridae | BDV | S (72, 208) | |||
Peribunyaviridae | PEAV* | T (46) | |||
Peribunyaviridae | SBV* | S+ (289, 549) | |||
Herpesviridae | BHV-1 | Penis, foreskin, U (518) | S (237) | ||
Retroviridae | BLV | S (162, 237) | |||
Retroviridae | BIV | S (495) | |||
Poxviridae | LSDV | T+, E+ (24) | S+ (316) | ||
Picornaviridae | FMDV | S (119) | |||
Reoviridae | BTV* | S (706) | R; E; T; A | ||
Sheep | Reoviridae | BTV* | T+, E, P, BG (endothelial cells) (560) | S+ (383) | E; A; T |
Phenuiviridae | RVFV*†a | T (endothelial cells, fibroblasts, smooth muscles, Mϕ) (514) | |||
Retroviridae | MVV | T, E, SV, P, BG, AG (543, 595) | S (595) | ||
Retroviridae | CAEV | T, E, VD, SV, P, BG (10, 699) | S (10) | ||
Retrovirida | SRLV genotype E | T+ (255) | |||
Horse | Herpesviridae | EHV | T+ (endothelial cells, macrophages), E, P+ (epithelial cells) (294, 677) | S (10, 279, 677, 721) | A; E |
Arteriviridae | EAV | T, E+, VD+, AG+d, P+, BG+, SV+ | S+d (87, 293, 684) | V; R |
Although ZIKV was the first arbovirus to be reported in human semen in 2015 (493), five arboviruses were already known to infect the semen of pigs [African swine fever virus (ASFV), classical swine fever virus (CSFV), JEV] and bulls [Schmallenberg virus (SBV) and bluetongue virus (BTV)]. This list now includes seminal excretion of Peaton virus (PEAV) in bulls, BTV and Rift Valley fever virus (RVFV) in sheep, and CHIKV, RVFV, and DENV in humans, suggesting that MGT infection by arboviruses is more common than expected (TABLE 8). Importantly, some of these animal arboviruses are zoonotic (JEV, RVFV) and impact embryonic development (BTV, CSFV) (TABLE 8). The attention paid to ZIKV in terms of sexual transmission and congenital effects, therefore, should be extended to other arboviruses that affect humans. Although the detection of a given zoonotic virus in the animal MGT does not imply that this virus will infect the same organs/cells in humans, it can reasonably be argued that it deserves investigation. As an example, RVFV, a zoonotic arbovirus that infects the sheep testis (514), was found in the semen of an immunosuppressed, infected patient 4 mo post-symptoms (277). Two other zoonotic viruses, JEV and HEV, infect the MGT of animals and are forecasted as future pathogens for North American populations (476).
In contrast to humans, the persistence in the MGT of acute viruses has long been recognized in farm animals, with over eight viruses from different families reported in semen, and in at least one male genital organ, in pigs, bulls, sheep, and horses (TABLE 8). A large panel of MGT organs and cells, including germ cells, was associated with virus persistence in animal semen (TABLE 8). The origin of the viruses persistently shed in semen appears complex. For example, the accumulation of data on porcine reproductive and respiratory syndrome virus (PRRSV)-infected animals suggests that seminal viruses arise from testicular germ cells during the acute phase of the infection, and from infected macrophages that infiltrate MGT organs during the persistent phase (554). Lifelong MGT infection and seminal shedding were demonstrated for two acute animal viruses, equine arteritis virus (EAV) in horses (37) and bovine viral diarrhea virus (BVDV) in bulls (75). Consequently, the infected male horses and bulls are major viral reservoirs, spreading the disease among cattle herds. The processes generating these lifelong infected animals differ between EAV and BVDV. EAV persistence is specific to the MGT, testosterone dependent, and concerns 10–70% of stallions infected during adulthood; the vas deferens ampullae was identified as the main virus reservoir (443). Long-term persistent infection has been correlated with the susceptibility to EAV infection of a subpopulation of ampullae-specific CD3+ T lymphocytes expressing an isoform of the CXCL16 gene with EAV receptor activity (37). The mechanisms responsible for immune evasion during the persistent infection of this MGT duct are still under investigation. In contrast to EAV, lifelong BVDV infection is not restricted to MGT and results from infection during fetal life, before the development of the immune system. The neonate calves become immunotolerant to BVDV, recognized as self, and remain viremic, shedding virus in semen lifelong (75, 542).
IX. CONCLUSIONS AND FUTURE DIRECTIONS
Viral infections of the MGT raise a broad range of issues at the interface between the fields of reproductive biology/endocrinology/urology (e.g., infertility, endocrine disturbances, genital cancers) and infectious diseases/virology (e.g., horizontal and vertical transmission of viruses, viral reservoirs, and persistence). Studies on viral infections in the MGT have significantly increased over the last two decades in these distinct fields. However, much remains to be done to improve our understanding of the interplay between viruses and the MGT and bring mechanistic insights to help targeted therapeutics interventions. We highlight below research questions and needs pertaining to these different issues and conclude with proposition of approaches.
A. Viral Infections and Infertility: What Are the Mechanisms at Play?
In addition to the well-known deleterious consequences of MuV on testis functions, an accumulation of observational data incriminate viral infections, especially chronic (e.g., hepatitis viruses, HPV), in altered semen/sperm parameters and diminished fertilization abilities. The mechanisms behind those alterations are largely unknown. To discriminate between indirect systemic (e.g., generalized inflammation, oxidative stress, endocrine disturbances at the hypothalamic-pituitary level) and direct negative effects of viral infections on MGT reproductive functions, there is now a need to deepen these data “at the bench.” The focus should not only be on ejaculated spermatozoa (the end product) but also on testis, epididymis, and accessory gland functions to establish whether the incriminated viruses modify the cellular homeostasis and male tissue microenvironment. Moreover, in the view of the paucity of data, further investigations on the hypothalomo-pituitary gonadal axis functions in men suffering from MGT infections are warranted. Importantly also, it is now established that environmental factors can affect epigenetic marks, and that nongenetic alterations can be passed on by the male gametes to the offspring (727). Whether testis and epididymis infections and/or an inflammatory environment can impact the gamete epigenome and in turn its ability to produce viable offspring has never been explored.
B. Viruses and Cancers of Male Genital Organs: Is There Anything Left to Uncover?
Whether viral infections play a part in the etiology or in the progression of prostate and testis cancers is unproven, despite numerous studies investigating the implication of well-known oncogenic viruses. There are most likely multiple environmental contributors to MGT cancers, which interact with each other and with host genetics and immune system in a temporal manner. Measuring the exposome of the individuals, defined as the totality of environmental exposures from conception onward including chemicals and infectious agents, should advance our understanding of environmental contributors. As shown in FIGURE 6, MGT organs, including the prostate and the testis, are the target of multiple viruses. Viral metagenomics recently revealed the existence of a human virome (i.e., viral community in tissues) that, as suspected for the microbiome, might influence health and disease (569). Viral metagenomics, as a wide unbiased approach, has great potential for uncovering novel viruses as well as yet unsuspected known viruses or specific viral combinations in diseased versus healthy MGT samples.
C. Viral Integration in the Human Germline and Endogenization: Can It Happen Again?
Genome sequencing and paleovirology have revealed the fascinating interactions between viruses and the gametes, demonstrating repetitive infections of the germline during evolution by both retroviruses and non-retroviral viruses, and highlighting the possibility of new viral integrations by a whole range of contemporary viruses. Whether HIV and non-retroviral viruses that associate with germ cells and spermatozoa could be “en route” for integration in the germline and future endogenization in humans is indeed an exciting question. Deciphering how human male germ cells are equipped to face viral colonization should provide some interesting answers.
D. Vertical Transmission of Viruses Through Semen: An Important Issue for Medically Assisted Reproduction
In the context of increasing use of medically assisted reproduction, it is crucial to determine whether the different semen compartments (i.e., seminal fluid, spermatozoa, and/or other semen cells) or the testicular germ cells (in the case of ICSI) are targeted by emerging viruses to prevent transmission to the partner and embryo. Indeed, virus-contaminated seminal fractions may induce congenital defects and/or viral chronicity (156, 218, 350). Long-lasting ZIKV infection in infants after in utero acquisition (74) and reports of fetal testis infection by ZIKV (45) justifies careful monitoring of early life infection cases.
E. Sexual Transmission of Viruses Through Semen: How Can We Anticipate Future Epidemics?
The transmission of viral diseases through semen is more than ever a major global health concern. Indeed, after ZIKV and EBOV, several emerging viruses forecasted to generate future outbreaks in human population (476, 492) represent suspects for sexual transmission (TABLE 7). As highlighted in this review, sexual transmission through semen represents a very powerful mode of dissemination for viruses and is extremely complicated to control, as exemplified by the HIV-1 pandemic and its over 35 million deaths, and by the resurgence of EBOV epidemics foci following sexual transmission. Whether the acquisition of sexual transmission mode by emerging viruses reflects virus evolution associated with novel tissue tropism or changes in host ecology is unknown (25). Inter- and intra-individual analysis of genomic sequences of sexually transmittable emerging viruses would be helpful to determine if unique signatures are involved in MGT tropism. Monitoring and controlling pathogens in non-human species including NHP and livestock is essential for anticipating human epidemics since most emerging viruses are zoonotic. This is underlined in the concept launch in 2004 of “one world-one health,” whereby animal and human health are considered inter-dependent (www.oneworldonehealth.org). Assessing MGT tropism of animal viruses in their host as well as in human tissues, for instance using ex vivo approaches, should greatly improve anticipating and in turn preventing sexual transmission. As highlighted in this review, a number of zoonotic viruses persist in animal semen (including HEV and JEV), suggesting potential for sexual transmission in future outbreaks (476, 718). Importantly however, molecular detection of viral sequences in semen is insufficient to determine capacity for sexual transmission. Sensitive culture techniques along with molecular detection of replication-competent virus in semen are warranted to assess the transmission potential of these viruses.
F. Mechanisms of Viral Persistence in Semen: The New Quest
Researches in HIV-infected patients over the last 30 yr, along with the recent reports of viral persistence in cured EBOV and ZIKV patients, have demonstrated that systemic viruses in semen do not exclusively arise from passive blood diffusion and have uncovered the existence of local MGT sources. Yet, we know very little about these sources and the specificities of their tissue microenvironment. Because of its status as an immune-privileged organ and as a pharmacological sanctuary, the testis is widely considered to be responsible for viral persistence in semen. However, recent data in HIV-infected patients and in ZIKV-infected men with undetectable viremia have demonstrated that other MGT organs can also be sites of persistent infection (30, 200, 440). Interestingly, we recently demonstrated that multiple MGT organs seed both virus and infected cells in the semen from SIV-infected monkeys and that these sources vary among individuals (299). These findings, which need to be confirmed in larger studies, echo the inter-individual variations of virus excretion pattern and persistence in semen (TABLE 7). Many questions await to be answered: What are the cellular and molecular actors that support viral persistence? What are the host/viral determinants leading to the continuous/intermittent excretion from specific MGT organs into the seminal lumen of virions and infected cells? Is this persistence the result of active replication or long-term maintenance of a stock of virus with little de novo viral production? How are viruses eventually cleared from the MGT?
G. Target Cells and Routes of Infection of MGT Organs by Systemic Viruses: Still a Lot to Learn
Our understanding of the routes of infection of both external and internal MGT organs and of viruses’cell targets in the MGT is surprisingly limited, even for viruses, such as HBV, for which sexual transmission represents a prime mode of dissemination. Recent data suggest that alternative routes to hematogenous spread exist that involve virus exchange between MGT organs (299, 696). Moreover, many viruses reported in the MGT are neurotropic (371, 405), and viral trafficking between the MGT and the central nervous system through nerves has been described (219–221, 668, 767) for several human viruses (371). Neuronal pathways might allow viruses to traffic between immune-privilege niches like the brain and testis and escape from immune surveillance potentially leading to virus persistence and cross-reservoir seeding. Since several emerging viruses have become neurotropic (e.g., ZIKV, WNV, DENV) or able to cause neurological diseases (EBOV) (490), it would be of great interest to establish whether these viruses can disseminate to the MGT through nerves.
H. Summary of Research Needs and Proposition of Approaches
Answering the questions raised in this review implies three prerequisites, namely, 1) decipher for each virus of interest the nature of infected cells in the different male genital organs. Even for viruses with an established tropism for the MGT, we know surprisingly very little about their targeted organs and cells. This is an obstacle for understanding their action mechanisms in the range of pathological processes described above. 2) Determine the immunological specificities of the different male genital organs in humans. While the basis for immune-privilege is well established in rodent testis, it has been poorly investigated in humans. In addition, recent findings revealed that the testis is not the only viral reservoir in the MGT. Thus attention needs to be paid as to how MGT organs other than testis can also shelter viruses from adaptive immune system. 3) Establish the species and cells innate immunity differences in MGT organs, along with the mobilization of innate effectors and repressors upon exposure to different viral pathogens. As pointed out in this review, evidence is building up of unique features for testicular germ cells as well as for human testis versus rodents.
To achieve this, there is a need for a combination of animal models and ex vivo models of human MGT tissues. The choice of animal model is however complicated by both the host specificity of the virus under investigation and the necessity for morphological and immunological features of MGT organs close to humans. As highlighted in this review, caution is needed when extrapolating from rodent models. NHPs constitute a model of choice for many human viruses (e.g., HIV) and their MGT is very similar to humans. Nevertheless, NHPs raise cost and ethical issues. Pigs, which share many testis features with humans and natural virus tropism for MGT, might represent an interesting alternative.
We have provided in this review several examples of the usefulness of ex vivo approaches for the study of human MGT organs’ infections. The development of three-dimensional bio-printing, organoids, and microfluidic systems hold great promise for long-term maintenance and improved reconstruction of human tissues. In addition to the preservation of primary cells characteristic in long-term culture, organoids obtained through the reorganization of different isolated cell types offer the extra bonus of possible genome editing, which could provide useful insights into host/virus interactions (18). However, while brain and digestive tract organoids are becoming more and more robust and standardized, there are currently no satisfactory examples of male genital organoids, especially testis. Indeed, the very complex testis structure (e.g., cell-to-cell contacts, specific spatial association of germ cells along the tubules. . .) is a major obstacle to faithful reconstruction. Microfluidic cultures of testis tissue may represent a better alternative; indeed, microfluidic culture of mouse testis explants succeeded in maintaining functional tissue for over 6 mo and allowed full spermatogenesis (365).
Finally, the onset of viral metagenomics (virome) as well as the incrementation of human organs and cells’ atlas should provide key elements to improve our understanding of the interplay between viruses and the MGT and inter-individuals variability.
To conclude, the many gaps in knowledge of the organs/cells and molecular processes involved in the infection and persistence of viruses in the MGT urgently need to be addressed if we want to develop tools to prevent sexual transmission, reproductive disorders, chronic inflammation, and cancer development/progression.
GRANTS
N. Dejucq-Rainsford has received funding from European Union's Horizon 2020 research and innovation program under grant agreement no. 733176, ZIKAPLAN (grant agreement no. 734584), and ZIKALLIANCE (grant agreement no. 734548); REACTing; ANRS; Sidaction; Région Bretagne; INSERM; French Ministry of Health; and Rennes 1 University. G. Matusali received support from Sidaction and REACTing, and L. Houzet was supported by Sidaction/Fond dotation P. Bergé, Région Bretagne, and ZIKALLIANCE.
DISCLOSURES
No conflicts of interest, financial or otherwise, are declared by the authors.
ACKNOWLEDGMENTS
A. Le Tortorec and G. Matusali contributed equally to the review.
Address for reprint requests and other correspondence: N. Dejucq-Rainsford, IRSET-Inserm U1085, 9 avenue du Pr Léon Bernard, F-35000 Rennes, France (e-mail: nathalie.
REFERENCES
- 1. . Polarized cell migration during cell-to-cell transmission of herpes simplex virus in human skin keratinocytes. J Virol 87: 7921–7932, 2013. doi:10.1128/JVI.01172-13.
Crossref | PubMed | Web of Science | Google Scholar - 2. . Potential Impact of Sexual Transmission on Ebola Virus Epidemiology: Sierra Leone as a Case Study. PLoS Negl Trop Dis 10:
e0004676 , 2016. doi:10.1371/journal.pntd.0004676.
Crossref | PubMed | Web of Science | Google Scholar - 3. . Viral etiology of prostate cancer: Genetic alterations and immune response. A literature review. Int J Surg 52: 136–140, 2018. doi:10.1016/j.ijsu.2018.02.050.
Crossref | PubMed | Web of Science | Google Scholar - 4. . Pituitary-testicular interrelationships in mumps orchitis and other viral infections. BMJ 1: 1177–1180, 1978. doi:10.1136/bmj.1.6121.1177.
Crossref | PubMed | Google Scholar - 5. . Zika virus persistence in the central nervous system and lymph nodes of Rhesus monkeys. Cell 169: 610–620.e14, 2017. doi:10.1016/j.cell.2017.04.008.
Crossref | PubMed | Web of Science | Google Scholar - 6. . Endogenous viruses: connecting recent and ancient viral evolution. Virology 479-480: 26–37, 2015. doi:10.1016/j.virol.2015.02.011.
Crossref | PubMed | Web of Science | Google Scholar - 7. . Androgen and estrogen production in elderly men with gynecomastia and testicular atrophy after mumps orchitis. J Clin Endocrinol Metab 50: 380–386, 1980. doi:10.1210/jcem-50-2-380.
Crossref | PubMed | Web of Science | Google Scholar - 8. . Male circumcision and genital human papillomavirus: a systematic review and meta-analysis. Sex Transm Dis 39: 104–113, 2012. doi:10.1097/OLQ.0b013e3182387abd.
Crossref | PubMed | Web of Science | Google Scholar - 9. . Male circumcision and the incidence and clearance of genital human papillomavirus (HPV) infection in men: the HPV Infection in men (HIM) cohort study. BMC Infect Dis 14: 75, 2014. doi:10.1186/1471-2334-14-75.
Crossref | PubMed | Web of Science | Google Scholar - 10. . Detection of viral genomes of caprine arthritis-encephalitis virus (CAEV) in semen and in genital tract tissues of male goat. Theriogenology 69: 473–480, 2008. doi:10.1016/j.theriogenology.2007.10.017.
Crossref | PubMed | Web of Science | Google Scholar - 11. . Expression of hepatitis B virus genes in early embryonic cells originated from hamster ova and human spermatozoa transfected with the complete viral genome. Asian J Androl 8: 273–279, 2006. doi:10.1111/j.1745-7262.2006.00141.x.
Crossref | PubMed | Web of Science | Google Scholar - 12. . Detection and expression of hepatitis B virus X gene in one and two-cell embryos from golden hamster oocytes in vitro fertilized with human spermatozoa carrying HBV DNA. Mol Reprod Dev 70: 30–36, 2005. doi:10.1002/mrd.20185.
Crossref | PubMed | Web of Science | Google Scholar - 13. . Cytomegalovirus infection of seminal vesicles. Int J Surg Pathol 24: 720, 2016. doi:10.1177/1066896916657592.
Crossref | PubMed | Web of Science | Google Scholar - 14. . The risk of sustained sexual transmission of Zika is underestimated. PLoS Pathog 13:
e1006633 , 2017. doi:10.1371/journal.ppat.1006633.
Crossref | PubMed | Web of Science | Google Scholar - 15. . Characterization of the influence of semen-derived enhancer of virus infection on the interaction of HIV-1 with female reproductive tract tissues. J Virol 89: 5569–5580, 2015. doi:10.1128/JVI.00309-15.
Crossref | PubMed | Web of Science | Google Scholar - 16. . Prostatitis and male infertility. J Reprod Immunol 100: 30–36, 2013. doi:10.1016/j.jri.2013.05.004.
Crossref | PubMed | Web of Science | Google Scholar - 17. . An update on HDV: virology, pathogenesis and treatment. Antivir Ther 18, 3 Pt B: 541–548, 2013. doi:10.3851/IMP2598.
Crossref | PubMed | Web of Science | Google Scholar - 18. . Testicular organoids: a new model to study the testicular microenvironment in vitro? Hum Reprod Update 24: 176–191, 2018. doi:10.1093/humupd/dmx036.
Crossref | PubMed | Web of Science | Google Scholar - 19. . HIV infection and immune defense of the penis. Am J Reprod Immunol 65: 220–229, 2011. doi:10.1111/j.1600-0897.2010.00941.x.
Crossref | PubMed | Web of Science | Google Scholar - 20. . Modeling mucosal cell-associated HIV type 1 transmission in vitro. J Infect Dis 210, Suppl 3: S648–S653, 2014. doi:10.1093/infdis/jiu537.
Crossref | PubMed | Web of Science | Google Scholar - 21. . Targeting trojan horse leukocytes for HIV prevention. AIDS 24: 163–187, 2010. doi:10.1097/QAD.0b013e32833424c8.
Crossref | PubMed | Web of Science | Google Scholar - 22. . Human male genital tract immunity and experimental models. In: Mucosal Immunology, edited by Mestecky J, Lamm ME, Strober W, Bienenstock J, McGhee JR, Mayer L. New York: Elsevier Academic, 2005, p. 1647–1659.
Crossref | Google Scholar - 23. . Genital HPV infection and related lesions in men. Prev Med 53, Suppl 1: S36–S41, 2011. doi:10.1016/j.ypmed.2011.08.002.
Crossref | PubMed | Web of Science | Google Scholar - 24. . Sites of persistence of lumpy skin disease virus in the genital tract of experimentally infected bulls. Reprod Domest Anim 45: 250–255, 2010. doi:10.1111/j.1439-0531.2008.01274.x.
Crossref | PubMed | Web of Science | Google Scholar - 25. . The evolution of transmission mode. Philos Trans R Soc Lond B Biol Sci 372:
20160083 , 2017. doi:10.1098/rstb.2016.0083.
Crossref | PubMed | Web of Science | Google Scholar - 26. . The latent human herpesvirus-6A genome specifically integrates in telomeres of human chromosomes in vivo and in vitro. Proc Natl Acad Sci USA 107: 5563–5568, 2010. doi:10.1073/pnas.0913586107.
Crossref | PubMed | Web of Science | Google Scholar - 27. . The immune privilege of testis and gravid uterus: same difference? Mol Cell Endocrinol 382: 509–520, 2014. doi:10.1016/j.mce.2013.09.022.
Crossref | PubMed | Web of Science | Google Scholar - 28. . Systemic distribution of West Nile virus infection: postmortem immunohistochemical study of six cases. Brain Pathol 17: 354–362, 2007. doi:10.1111/j.1750-3639.2007.00080.x.
Crossref | PubMed | Web of Science | Google Scholar - 29. . Naturally occurring fragments from two distinct regions of the prostatic acid phosphatase form amyloidogenic enhancers of HIV infection. J Virol 86: 1244–1249, 2012. doi:10.1128/JVI.06121-11.
Crossref | PubMed | Web of Science | Google Scholar - 30. . Probable sexual transmission of Zika virus from a vasectomised man. Lancet Infect Dis 16: 1107, 2016. doi:10.1016/S1473-3099(16)30320-6.
Crossref | PubMed | Web of Science | Google Scholar - 31. . Persistence of Nipah virus RNA in semen of survivor. Clin Infect Dis 69: 377–378, 2019. doi:10.1093/cid/ciy1092.
Crossref | PubMed | Web of Science | Google Scholar - 32. . Paleovirology and virally derived immunity. Trends Ecol Evol 27: 627–636, 2012. doi:10.1016/j.tree.2012.07.007.
Crossref | PubMed | Web of Science | Google Scholar - 33. . Surface plasmon resonance analysis of gp17, a natural CD4 ligand from human seminal plasma inhibiting human immunodeficiency virus type-1 gp120-mediated syncytium formation. Eur J Biochem 245: 208–213, 1997. doi:10.1111/j.1432-1033.1997.00208.x.
Crossref | PubMed | Google Scholar - 34. . Randomized, controlled intervention trial of male circumcision for reduction of HIV infection risk: the ANRS 1265 Trial. [Correction in PLoS Med 3: e226, 2006.] PLoS Med 2:
e298 , 2005. doi:10.1371/journal.pmed.0020298.
Crossref | PubMed | Web of Science | Google Scholar - 35. . Human papillomavirus 16 infection as a potential risk factor for prostate cancer: an adaptive meta-analysis. Epidemiol Health 37:
e2015005 , 2015. doi:10.4178/epih/e2015005.
Crossref | PubMed | Web of Science | Google Scholar - 36. . Male circumcision for HIV prevention in young men in Kisumu, Kenya: a randomised controlled trial. Lancet 369: 643–656, 2007. doi:10.1016/S0140-6736(07)60312-2.
Crossref | PubMed | Web of Science | Google Scholar - 37. . Reproductive effects of arteriviruses: equine arteritis virus and porcine reproductive and respiratory syndrome virus infections. Curr Opin Virol 27: 57–70, 2017. doi:10.1016/j.coviro.2017.11.005.
Crossref | PubMed | Web of Science | Google Scholar - 38. . Prolonged shedding of Chikungunya virus in semen and urine: a new perspective for diagnosis and implications for transmission. IDCases 6: 100–103, 2016. doi:10.1016/j.idcr.2016.10.007.
Crossref | PubMed | Web of Science | Google Scholar - 39. . The evolutionary dynamics of human endogenous retroviral families. Annu Rev Genomics Hum Genet 7: 149–173, 2006. doi:10.1146/annurev.genom.7.080505.115700.
Crossref | PubMed | Web of Science | Google Scholar - 40. . Yellow fever virus RNA in urine and semen of convalescent patient, Brazil. Emerg Infect Dis 24: 176–178, 2018. doi:10.3201/eid2401.171310.
Crossref | PubMed | Web of Science | Google Scholar - 41. . Evidence of Ebola virus replication and high concentration in semen of a patient during recovery. Clin Infect Dis 65: 1400–1403, 2017. doi:10.1093/cid/cix518.
Crossref | PubMed | Web of Science | Google Scholar - 42. . Enhancement of Ebola virus infection by seminal amyloid fibrils. Proc Natl Acad Sci USA 115: 7410–7415, 2018. doi:10.1073/pnas.1721646115.
Crossref | PubMed | Web of Science | Google Scholar - 43. . Virus and antibody dynamics in travelers with acute Zika virus infection. Clin Infect Dis 66: 1173–1180, 2018. doi:10.1093/cid/cix967.
Crossref | PubMed | Web of Science | Google Scholar - 44. . Demonstration of developmental anomalies in mouse fetuses by transfer of murine cytomegalovirus DNA-injected eggs to surrogate mothers. J Infect Dis 167: 1288–1295, 1993. doi:10.1093/infdis/167.6.1288.
Crossref | PubMed | Web of Science | Google Scholar - 45. . A clinical and histopathological study of malformations observed in fetuses infected by the Zika virus. Brain Pathol 29: 114–125, 2019. doi:10.1111/bpa.12644.
Crossref | PubMed | Web of Science | Google Scholar - 46. . First genomic detection of Peaton virus in a calf with hydranencephaly in Israel. Vet Med Sci 5: 87–92, 2019. doi:10.1002/vms3.129.
Crossref | PubMed | Web of Science | Google Scholar - 47. . A role for CD95 ligand in preventing graft rejection. Nature 377: 630–632, 1995. doi:10.1038/377630a0.
Crossref | PubMed | Web of Science | Google Scholar - 48. . Long-term reinfection of the human genome by endogenous retroviruses. Proc Natl Acad Sci USA 101: 4894–4899, 2004. doi:10.1073/pnas.0307800101.
Crossref | PubMed | Web of Science | Google Scholar - 49. . Unexpected inheritance: multiple integrations of ancient bornavirus and ebolavirus/marburgvirus sequences in vertebrate genomes. PLoS Pathog 6:
e1001030 , 2010. doi:10.1371/journal.ppat.1001030.
Crossref | PubMed | Web of Science | Google Scholar - 50. . Seminal plasma promotes the attraction of Langerhans cells via the secretion of CCL20 by vaginal epithelial cells: involvement in the sexual transmission of HIV. Hum Reprod 21: 1135–1142, 2006. doi:10.1093/humrep/dei496.
Crossref | PubMed | Web of Science | Google Scholar - 51. . The immune tolerant phase of chronic HBV infection: new perspectives on an old concept. Cell Mol Immunol 12: 258–263, 2015. doi:10.1038/cmi.2014.79.
Crossref | PubMed | Web of Science | Google Scholar - 52. . Endogenous retrovirus drives hitherto unknown proapoptotic p63 isoforms in the male germ line of humans and great apes. Proc Natl Acad Sci USA 108: 3624–3629, 2011. doi:10.1073/pnas.1016201108.
Crossref | PubMed | Web of Science | Google Scholar - 53. . Prevalence of sexually transmissible pathogens in semen from asymptomatic male infertility patients with and without leukocytospermia. Fertil Steril 87: 1087–1097, 2007. doi:10.1016/j.fertnstert.2006.08.109.
Crossref | PubMed | Web of Science | Google Scholar - 54. . Prevalence of human herpesvirus types 1-8 in the semen of infertility patients and correlation with semen parameters. Fertil Steril 76: 416–418, 2001. doi:10.1016/S0015-0282(01)01920-3.
Crossref | PubMed | Web of Science | Google Scholar - 55. . The macrophages in testis function. J Reprod Immunol 119: 107–112, 2017. doi:10.1016/j.jri.2016.06.008.
Crossref | PubMed | Web of Science | Google Scholar - 56. . Differential activation of inflammatory pathways in testicular macrophages provides a rationale for their subdued inflammatory capacity. J Immunol 194: 5455–5464, 2015. doi:10.4049/jimmunol.1401132.
Crossref | PubMed | Web of Science | Google Scholar - 57. . Humanized mouse model of Ebola virus disease mimics the immune responses in human disease. J Infect Dis 213: 703–711, 2016. doi:10.1093/infdis/jiv538.
Crossref | PubMed | Web of Science | Google Scholar - 58. . Vertical transmission of HTLV-I/II: a review. Rev Inst Med Trop São Paulo 40: 245–251, 1998. doi:10.1590/S0036-46651998000400008.
Crossref | PubMed | Google Scholar - 59. . Mumps virus recovered from testicles by fine-needle aspiration biopsy in cases of mumps orchitis. Scand J Infect Dis 5: 3–5, 1973. doi:10.3109/inf.1973.5.issue-1.02.
Crossref | PubMed | Google Scholar - 60. . Persistent Viral Infections (4th ed.). Galveston, TX: Univ. of Texas Medical Branch at Galveston, 1996.
Google Scholar - 61. . Clinical and virological characterization of persistent human infection with simian foamy viruses. AIDS Res Hum Retroviruses 23: 1330–1337, 2007. doi:10.1089/aid.2007.0104.
Crossref | PubMed | Web of Science | Google Scholar - 62. . Ebola virus infection associated with transmission from survivors. Emerg Infect Dis 25: 249–255, 2019. doi:10.3201/eid2502.181011.
Crossref | PubMed | Web of Science | Google Scholar - 63. . Testicular hypoplasia in a bull persistently infected with bovine diarrhoea virus. J Comp Pathol 137: 169–173, 2007. doi:10.1016/j.jcpa.2007.06.002.
Crossref | PubMed | Web of Science | Google Scholar - 64. . Herpes simplex virus infection of human dendritic cells induces apoptosis and allows cross-presentation via uninfected dendritic cells. J Immunol 174: 2220–2227, 2005. doi:10.4049/jimmunol.174.4.2220.
Crossref | PubMed | Web of Science | Google Scholar - 65. . Intraepithelial and stromal lymphocytes in the normal human prostate. Prostate 55: 187–193, 2003. doi:10.1002/pros.10224.
Crossref | PubMed | Web of Science | Google Scholar - 66. . Opsonization of HIV-1 by semen complement enhances infection of human epithelial cells. J Immunol 169: 3301–3306, 2002. doi:10.4049/jimmunol.169.6.3301.
Crossref | PubMed | Web of Science | Google Scholar - 67. . Prospective evaluation of the threat related to the use of seminal fractions from hepatitis C virus-infected men in assisted reproductive techniques. Hum Reprod 24: 530–535, 2009. doi:10.1093/humrep/den414.
Crossref | PubMed | Web of Science | Google Scholar - 68. . Vaginal pH neutralization by semen as a cofactor of HIV transmission. Clin Microbiol Infect 3: 19–23, 1997. doi:10.1111/j.1469-0691.1997.tb00246.x.
Crossref | PubMed | Google Scholar - 69. . Heparin-binding domain of human fibronectin binds HIV-1 gp120/160 and reduces virus infectivity. J Med Virol 54: 44–53, 1998. doi:10.1002/(SICI)1096-9071(199801)54:1<44:AID-JMV7>3.0.CO;2-P.
Crossref | PubMed | Web of Science | Google Scholar - 70. . Etiologies of nongonococcal urethritis: bacteria, viruses, and the association with orogenital exposure. J Infect Dis 193: 336–345, 2006. doi:10.1086/499434.
Crossref | PubMed | Web of Science | Google Scholar - 71. . A comparison of seminal hepatitis C virus (HCV) RNA levels during recent and chronic HCV infection in HIV-infected and HIV-uninfected individuals. J Infect Dis 211: 736–743, 2015. doi:10.1093/infdis/jiu550.
Crossref | PubMed | Web of Science | Google Scholar - 72. . Short communication: transmission of border disease virus to seronegative cows inseminated with infected semen. Res Vet Sci 100: 297–298, 2015. doi:10.1016/j.rvsc.2015.03.027.
Crossref | PubMed | Web of Science | Google Scholar - 73. . Hepatitis C virus in the semen of men coinfected with HIV-1: prevalence and origin. AIDS 19: 1827–1835, 2005. doi:10.1097/01.aids.0000189847.98569.2d.
Crossref | PubMed | Web of Science | Google Scholar - 74. . Persistent detection of Zika virus RNA from an infant with severe microcephaly - a case report. BMC Infect Dis 18: 388, 2018. doi:10.1186/s12879-018-3313-4.
Crossref | PubMed | Web of Science | Google Scholar - 75. . Bovine viral diarrhea virus infections: manifestations of infection and recent advances in understanding pathogenesis and control. Vet Pathol 51: 453–464, 2014. doi:10.1177/0300985813520250.
Crossref | PubMed | Web of Science | Google Scholar - 76. . The effect of a single sauna exposure on spermatozoa. Arch Androl 12: 9–15, 1984. doi:10.3109/01485018409161141.
Crossref | PubMed | Google Scholar - 77. . Region-specific innate antiviral responses of the human epididymis. Mol Cell Endocrinol 473: 72–78, 2018. doi:10.1016/j.mce.2018.01.004.
Crossref | PubMed | Web of Science | Google Scholar - 78. . Expression profiles of human epididymis epithelial cells reveal the functional diversity of caput, corpus and cauda regions. Mol Hum Reprod 22: 69–82, 2016. doi:10.1093/molehr/gav066.
Crossref | PubMed | Web of Science | Google Scholar - 79. . Common skin disorders of the penis. BJU Int 90: 498–506, 2002. doi:10.1046/j.1464-410X.2002.02962.x.
Crossref | PubMed | Web of Science | Google Scholar - 80. . Factors of intermittent HIV-1 excretion in semen and efficiency of sperm processing in obtaining spermatozoa without HIV-1 genomes. AIDS 18: 757–766, 2004. doi:10.1097/00002030-200403260-00006.
Crossref | PubMed | Web of Science | Google Scholar - 81. . HIV and mucosal barrier interactions: consequences for transmission and pathogenesis. Curr Opin Immunol 36: 22–30, 2015. doi:10.1016/j.coi.2015.06.004.
Crossref | PubMed | Web of Science | Google Scholar - 82. . Hepatitis B infection is highly endemic in Uganda: findings from a national serosurvey. Afr Health Sci 9: 98–108, 2009.
PubMed | Web of Science | Google Scholar - 83. . Interleukin-1 inhibits Leydig cell steroidogenesis in primary culture. Endocrinology 123: 1605–1610, 1988. doi:10.1210/endo-123-3-1605.
Crossref | PubMed | Web of Science | Google Scholar - 84. . Comparison of the effect of semen from HIV-infected and uninfected men on CD4+ T-cell infection. AIDS 30: 1197–1208, 2016. doi:10.1097/QAD.0000000000001048.
Crossref | PubMed | Web of Science | Google Scholar - 85. . Analysis of persistence of human papillomavirus infection in men evaluated by sampling multiple genital sites. Eur Rev Med Pharmacol Sci 19: 4153–4163, 2015.
PubMed | Web of Science | Google Scholar - 86. . Mucosal and systemic HIV-1-specific immunity in HIV-1-exposed but uninfected heterosexual men. AIDS 17: 531–539, 2003. doi:10.1097/00002030-200303070-00008.
Crossref | PubMed | Web of Science | Google Scholar - 87. . Equine arteritis virus has specific tropism for stromal cells and CD8+ T and CD21+ B lymphocytes but not for glandular epithelium at the primary site of persistent infection in the stallion reproductive tract. J Virol 91:
e00418-17 , 2017. doi:10.1128/JVI.00418-17.
Crossref | PubMed | Web of Science | Google Scholar - 88. . Spermatozoa capture HIV-1 through heparan sulfate and efficiently transmit the virus to dendritic cells. J Exp Med 206: 2717–2733, 2009. doi:10.1084/jem.20091579.
Crossref | PubMed | Web of Science | Google Scholar - 89. . The making of “transgenic spermatozoa”. Biol Reprod 68: 1477–1483, 2003. doi:10.1095/biolreprod.102.009340.
Crossref | PubMed | Web of Science | Google Scholar - 90. . STD Facts HPV [Online]. 2019. https://www.cdc.gov/std/hpv/stdfact-hpv.htm.
Google Scholar - 91. . STD Fact herpes [Online]. 2019. https://www.cdc.gov/std/herpes/stdfact-herpes.htm.
Google Scholar - 92. . Molluscum contagiosum Risk Factors [Online]. 2019. https://www.cdc.gov/poxvirus/molluscum-contagiosum/risk.html.
Google Scholar - 93. . Herpesvirus type 2 in the male genitourinary tract. Science 178: 318–319, 1972. doi:10.1126/science.178.4058.318.
Crossref | PubMed | Web of Science | Google Scholar - 95. . In vitro transformation by HSV-2 from a human prostatic carcinoma. IARC Sci Publ 11: 195–197, 1975.
PubMed | Google Scholar - 96. . Zika virus infection in dexamethasone-immunosuppressed mice demonstrating disseminated infection with multi-organ involvement including orchitis effectively treated by recombinant type i interferons. EBioMedicine 14: 112–122, 2016. doi:10.1016/j.ebiom.2016.11.017.
Crossref | PubMed | Web of Science | Google Scholar - 97. . Arboviruses and pregnancy: maternal, fetal, and neonatal effects. Lancet Child Adolesc Health 1: 134–146, 2017. doi:10.1016/S2352-4642(17)30021-4.
Crossref | PubMed | Web of Science | Google Scholar - 98. . Evidences of the inflammasome pathway in chronic prostatitis and chronic pelvic pain syndrome in an animal model. Prostate 73: 391–397, 2013. doi:10.1002/pros.22580.
Crossref | PubMed | Web of Science | Google Scholar - 99. . Seminal plasma induces global transcriptomic changes associated with cell migration, proliferation and viability in endometrial epithelial cells and stromal fibroblasts. Hum Reprod 29: 1255–1270, 2014. doi:10.1093/humrep/deu047.
Crossref | PubMed | Web of Science | Google Scholar - 100. . Testicular immunoregulation and spermatogenesis. Semin Cell Dev Biol 59: 157–165, 2016. doi:10.1016/j.semcdb.2016.01.019.
Crossref | PubMed | Web of Science | Google Scholar - 101. . Identification of pathogen signatures in prostate cancer using RNA-seq. PLoS One 10:
e0128955 , 2015. doi:10.1371/journal.pone.0128955.
Crossref | PubMed | Web of Science | Google Scholar - 102. . Niche adaptation and viral transmission of human papillomaviruses from archaic hominins to modern humans. PLoS Pathog 14:
e1007352 , 2018. doi:10.1371/journal.ppat.1007352.
Crossref | PubMed | Web of Science | Google Scholar - 103. . Interaction of programmed death-1 and programmed death-1 ligand-1 contributes to testicular immune privilege. Transplantation 87: 1778–1786, 2009. doi:10.1097/TP.0b013e3181a75633.
Crossref | PubMed | Web of Science | Google Scholar - 104. . Impact of early cART on HIV blood and semen compartments at the time of primary infection. PLoS One 12:
e0180191 , 2017. doi:10.1371/journal.pone.0180191.
Crossref | PubMed | Web of Science | Google Scholar - 105. . The estimated lifetime probability of acquiring human papillomavirus in the United States. Sex Transm Dis 41: 660–664, 2014. doi:10.1097/OLQ.0000000000000193.
Crossref | PubMed | Web of Science | Google Scholar - 106. . Latent sites of Epstein-Barr virus infection. Am J Clin Pathol 100: 502–506, 1993. doi:10.1093/ajcp/100.5.502.
Crossref | PubMed | Web of Science | Google Scholar - 107. . Localization of classical swine fever virus in male gonads during subclinical infection. J Gen Virol 83: 2717–2721, 2002. doi:10.1099/0022-1317-83-11-2717.
Crossref | PubMed | Web of Science | Google Scholar - 108. . Detection of classical swine fever virus in boar semen by reverse transcription-polymerase chain reaction. J Vet Diagn Invest 15: 35–41, 2003. doi:10.1177/104063870301500108.
Crossref | PubMed | Web of Science | Google Scholar - 109. . Persistence of porcine reproductive and respiratory syndrome virus in serum and semen of adult boars. J Vet Diagn Invest 7: 456–464, 1995. doi:10.1177/104063879500700406.
Crossref | PubMed | Web of Science | Google Scholar - 110. . Comparative histopathologic lesions of the male reproductive tract during acute infection of Zika virus in AG129 and Ifnar-/- mice. Am J Pathol 188: 904–915, 2018. doi:10.1016/j.ajpath.2017.12.019.
Crossref | PubMed | Web of Science | Google Scholar - 111. . Persistent Marburg virus infection in the testes of nonhuman primate survivors. Cell Host Microbe 24: 405–416.e3, 2018. doi:10.1016/j.chom.2018.08.003.
Crossref | PubMed | Web of Science | Google Scholar - 112. . Association between the JC polyomavirus infection and male infertility. PLoS One 7:
e42880 , 2012. doi:10.1371/journal.pone.0042880.
Crossref | PubMed | Web of Science | Google Scholar - 113. . Pathogenesis of experimental Ebola virus infection in guinea pigs. J Infect Dis 179, Suppl 1: S203–S217, 1999. doi:10.1086/514305.
Crossref | PubMed | Web of Science | Google Scholar - 114. . Lower genitourinary tract sources of seminal HIV. J Acquir Immune Defic Syndr 41: 430–438, 2006. doi:10.1097/01.qai.0000209895.82255.08.
Crossref | PubMed | Web of Science | Google Scholar - 115. . Association between culturable human immunodeficiency virus type 1 (HIV-1) in semen and HIV-1 RNA levels in semen and blood: evidence for compartmentalization of HIV-1 between semen and blood. J Infect Dis 177: 320–330, 1998. doi:10.1086/514213.
Crossref | PubMed | Web of Science | Google Scholar - 116. . Histology, immunohistochemistry, and in situ hybridization reveal overlooked Ebola virus target tissues in the Ebola virus disease guinea pig model. [Correction in Sci Rep 9: 15013, 2019.] Sci Rep 8: 1250, 2018. doi:10.1038/s41598-018-19638-x.
Crossref | PubMed | Web of Science | Google Scholar - 117. . New insights Into Marburg virus disease pathogenesis in the Rhesus Macaque model. J Infect Dis 218, Suppl_5: S423–S433, 2018. doi:10.1093/infdis/jiy367.
Crossref | PubMed | Web of Science | Google Scholar - 118. . The presence of human papillomavirus in semen does not affect the integrity of sperm DNA. Andrologia 49:
e12774 , 2017. doi:10.1111/and.12774.
Crossref | Web of Science | Google Scholar - 119. . Foot-and-mouth disease virus in semen of bulls and its transmission by artificial insemination. Arch Gesamte Virusforsch 23: 362–377, 1968. doi:10.1007/BF01242132.
Crossref | PubMed | Google Scholar - 120. . Role of semen on vaginal HIV-1 transmission and Maraviroc protection. Antimicrob Agents Chemother 59: 7847–7851, 2015. doi:10.1128/AAC.01496-15.
Crossref | PubMed | Web of Science | Google Scholar - 121. . Zika virus - reigniting the TORCH. Nat Rev Microbiol 14: 707–715, 2016. doi:10.1038/nrmicro.2016.125.
Crossref | PubMed | Web of Science | Google Scholar - 122. . Orchitis due to Coxsackie virus group B, type 5. Report of a case with isolation of virus from the testis. N Engl J Med 267: 498–500, 1962. doi:10.1056/NEJM196209062671008.
Crossref | PubMed | Web of Science | Google Scholar - 123. . Persistent HIV type 1 infection in semen and blood compartments in patients after long-term potent antiretroviral therapy. AIDS Res Hum Retroviruses 20: 1196–1209, 2004. doi:10.1089/aid.2004.20.1196.
Crossref | PubMed | Web of Science | Google Scholar - 124. . Complement OPSONIZATION promotes herpes simplex virus 2 infection of human dendritic cells. J Virol 90: 4939–4950, 2016. doi:10.1128/JVI.00224-16.
Crossref | PubMed | Web of Science | Google Scholar - 125. . Endogenous retrovirus expression in testis and epididymis. Biochem Soc Trans 35: 629–633, 2007. doi:10.1042/BST0350629.
Crossref | PubMed | Web of Science | Google Scholar - 126. . Detection, prevalence and analysis of emerging porcine parvovirus infections. Arch Virol 157: 1003–1010, 2012. doi:10.1007/s00705-012-1257-3.
Crossref | PubMed | Web of Science | Google Scholar - 127. . Virus-host studies in human seminal and mouse testicular cells. Acta Chir Hung 32: 83–90, 1991.
PubMed | Google Scholar - 128. . Testicular FasL is expressed by sperm cells. Proc Natl Acad Sci USA 98: 3316–3321, 2001. doi:10.1073/pnas.051566098.
Crossref | PubMed | Web of Science | Google Scholar - 129. . Evidence of sexual transmission of Zika virus. N Engl J Med 374: 2195–2198, 2016. doi:10.1056/NEJMc1604449.
Crossref | PubMed | Web of Science | Google Scholar - 130. . Inheritance of chromosomally integrated human herpesvirus 6 DNA. Blood 94: 1545–1549, 1999. doi:10.1182/blood.V94.5.1545.
Crossref | PubMed | Web of Science | Google Scholar - 131. . Poxviridae. In: Fields Virology, edited by Knipe DM, Howley PM. Philadelphia, PA: Lippincott Williams & Wilkins, 2013.
Google Scholar - 132. . Detection of hepatitis B virus DNA in spermatozoa, urine, saliva and leucocytes, of chronic HBsAg carriers. A lack of relationship with serum markers of replication. J Hepatol 4: 37–44, 1987. doi:10.1016/S0168-8278(87)80007-7.
Crossref | PubMed | Web of Science | Google Scholar - 133. . Ebola RNA persistence in semen of Ebola virus disease survivors - final report. N Engl J Med 377: 1428–1437, 2017. doi:10.1056/NEJMoa1511410.
Crossref | PubMed | Web of Science | Google Scholar - 134. . Macrophages Contribute to the Spermatogonial Niche in the Adult Testis. Cell Rep 12: 1107–1119, 2015. doi:10.1016/j.celrep.2015.07.015.
Crossref | PubMed | Web of Science | Google Scholar - 135. . Viruses in semen and male genital tissues--consequences for the reproductive system and therapeutic perspectives. Curr Pharm Des 10: 557–575, 2004. doi:10.2174/1381612043453225.
Crossref | PubMed | Web of Science | Google Scholar - 136. . The testicular antiviral defense system: localization, expression, and regulation of 2′5′ oligoadenylate synthetase, double-stranded RNA-activated protein kinase, and Mx proteins in the rat seminiferous tubule. J Cell Biol 139: 865–873, 1997. doi:10.1083/jcb.139.4.865.
Crossref | PubMed | Web of Science | Google Scholar - 137. . Interferon-alpha and -gamma expression in the rat testis. Endocrinology 136: 4925–4931, 1995. doi:10.1210/endo.136.11.7588226.
Crossref | PubMed | Web of Science | Google Scholar - 138. . Viruses in the mammalian male genital tract and their effects on the reproductive system. Microbiol Mol Biol Rev 65: 208–231, 2001. doi:10.1128/MMBR.65.2.208-231.2001.
Crossref | PubMed | Web of Science | Google Scholar - 139. . Expression of interferons-alpha and -gamma in testicular interstitial tissue and spermatogonia of the rat. Endocrinology 139: 3081–3087, 1998. doi:10.1210/endo.139.7.6083.
Crossref | PubMed | Web of Science | Google Scholar - 140. . Interferons and interferon-induced antiviral proteins in the testis. J Reprod Immunol 41: 291–300, 1998. doi:10.1016/S0165-0378(98)00065-5.
Crossref | PubMed | Web of Science | Google Scholar - 141. . Human immunodeficiency virus infects human seminal vesicles in vitro and in vivo. Am J Pathol 179: 2397–2408, 2011. doi:10.1016/j.ajpath.2011.08.005.
Crossref | PubMed | Web of Science | Google Scholar - 142. . The action of prostaglandin E2 on the human cervix: stimulation of interleukin 8 and inhibition of secretory leukocyte protease inhibitor. Am J Obstet Gynecol 180: 614–620, 1999. doi:10.1016/S0002-9378(99)70263-2.
Crossref | PubMed | Web of Science | Google Scholar - 143. . Herpesvirus type 2: isolation from seminal vesicle and testes. Urology 7: 541–544, 1976. doi:10.1016/0090-4295(76)90204-1.
Crossref | PubMed | Google Scholar - 144. . Resurgence of Ebola virus disease in Guinea linked to a survivor with virus persistence in seminal fluid for more than 500 days. Clin Infect Dis 63: 1353–1356, 2016. doi:10.1093/cid/ciw601.
Crossref | PubMed | Web of Science | Google Scholar - 145. . The Sertoli cell as the orchestra conductor of spermatogenesis: spermatogenic cells dance to the tune of testosterone. Hormones (Athens) 14: 479–503, 2015. doi:10.14310/horm.2002.1633.
Crossref | PubMed | Google Scholar - 146. . Visualization of HIV-1 interactions with penile and foreskin epithelia: clues for female-to-male HIV transmission. PLoS Pathog 11:
e1004729 , 2015. doi:10.1371/journal.ppat.1004729.
Crossref | PubMed | Web of Science | Google Scholar - 147. . The role of human papilloma virus in penile carcinogenesis and preneoplastic lesions: a potential target for vaccination and treatment strategies. Urol Clin North Am 43: 419–425, 2016. doi:10.1016/j.ucl.2016.06.003.
Crossref | PubMed | Web of Science | Google Scholar - 148. . No SEVI-mediated enhancement of rectal HIV-1 transmission of HIV-1 in two humanized mouse cohorts. Virology 488: 88–95, 2016. doi:10.1016/j.virol.2015.11.005.
Crossref | PubMed | Web of Science | Google Scholar - 149. . Parvovirus B19 is associated with benign testes as well as testicular germ cell tumours. Mol Pathol 52: 349–352, 1999. doi:10.1136/mp.52.6.349.
Crossref | PubMed | Google Scholar - 150. . Inflammatory-associated obstructions of the male reproductive tract. Andrologia 35: 321–324, 2003. doi:10.1111/j.1439-0272.2003.tb00866.x.
Crossref | PubMed | Web of Science | Google Scholar - 151. . Lineage-dependent differences in the disease progression of Zika virus infection in type-I interferon receptor knockout (A129) mice. PLoS Negl Trop Dis 11:
e0005704 , 2017. doi:10.1371/journal.pntd.0005704.
Crossref | PubMed | Web of Science | Google Scholar - 152. . Human endogenous retrovirus K and cancer: innocent bystander or tumorigenic accomplice? Int J Cancer 137: 1249–1257, 2015. doi:10.1002/ijc.29003.
Crossref | PubMed | Web of Science | Google Scholar - 153. . Characterisation of dendritic cell subsets in chronically inflamed human epididymis. Andrologia 48: 431–440, 2016. doi:10.1111/and.12463.
Crossref | PubMed | Web of Science | Google Scholar - 154. . Semen parameters during ZIKA virus infection in the olive baboon (Papio anubis). Fertil Steril 108: e52–e53, 2017. doi:10.1016/j.fertnstert.2017.07.168.
Crossref | Web of Science | Google Scholar - 155. . Evolutionary conflicts between viruses and restriction factors shape immunity. Nat Rev Immunol 12: 687–695, 2012. doi:10.1038/nri3295.
Crossref | PubMed | Web of Science | Google Scholar - 156. . Sexual transmission of Zika virus enhances in utero transmission in a mouse model. Sci Rep 8: 4510, 2018. doi:10.1038/s41598-018-22840-6.
Crossref | PubMed | Web of Science | Google Scholar - 157. . Frequent Zika virus sexual transmission and prolonged viral RNA shedding in an immunodeficient mouse model. Cell Rep 18: 1751–1760, 2017. doi:10.1016/j.celrep.2017.01.056.
Crossref | PubMed | Web of Science | Google Scholar - 158. . Prevalence of HPV infection among men: a systematic review of the literature. J Infect Dis 194: 1044–1057, 2006. doi:10.1086/507432.
Crossref | PubMed | Web of Science | Google Scholar - 159. . From ancestral infectious retroviruses to bona fide cellular genes: role of the captured syncytins in placentation. Placenta 33: 663–671, 2012. doi:10.1016/j.placenta.2012.05.005.
Crossref | PubMed | Web of Science | Google Scholar - 160. . A pair of co-opted retroviral envelope syncytin genes is required for formation of the two-layered murine placental syncytiotrophoblast. Proc Natl Acad Sci USA 108: E1164–E1173, 2011. doi:10.1073/pnas.1112304108.
Crossref | PubMed | Web of Science | Google Scholar - 161. . Alterations of seminal and hormonal parameters: an extrahepatic manifestation of HCV infection? World J Gastroenterol 12: 3073–3076, 2006. doi:10.3748/wjg.v12.i19.3073.
Crossref | PubMed | Web of Science | Google Scholar - 162. . Development of a PCR to diagnose BLV genome in frozen semen samples. Vet Microbiol 119: 10–18, 2007. doi:10.1016/j.vetmic.2006.08.030.
Crossref | PubMed | Web of Science | Google Scholar - 163. . Cytomegalovirus (CMV) infection--related to male and/or female infertility factors? Fertil Steril 91: 67–82, 2009. doi:10.1016/j.fertnstert.2007.11.014.
Crossref | PubMed | Web of Science | Google Scholar - 164. . Internalization of cystatin C in human cell lines. FEBS J 275: 4571–4582, 2008. doi:10.1111/j.1742-4658.2008.06600.x.
Crossref | PubMed | Google Scholar - 165. . Lymphocyte sub-populations in the male genital tract. Br J Urol 57: 769–774, 1985. doi:10.1111/j.1464-410X.1985.tb07051.x.
Crossref | PubMed | Google Scholar - 166. . Exosomes in HIV infection: a review and critical look. Infect Genet Evol 53: 146–154, 2017. doi:10.1016/j.meegid.2017.05.021.
Crossref | PubMed | Web of Science | Google Scholar - 167. . A case of Ebola virus infection. BMJ 2: 541–544, 1977. doi:10.1136/bmj.2.6086.541.
Crossref | PubMed | Google Scholar - 168. . Dermatoses of the glans penis and prepuce. J Am Acad Dermatol 37: 1–26, 1997. doi:10.1016/S0190-9622(97)70207-X.
Crossref | PubMed | Web of Science | Google Scholar - 169. . Zika virus and reproduction: facts, questions and current management. Hum Reprod Update 23: 629–645, 2017. doi:10.1093/humupd/dmx024.
Crossref | PubMed | Web of Science | Google Scholar - 170. . Testicular persistence of Parvovirus B19: evidence for preferential infection of germ cell tumors. Pathol Res Pract 204: 649–653, 2008. doi:10.1016/j.prp.2008.04.004.
Crossref | PubMed | Web of Science | Google Scholar - 171. . DNA of adeno-associated virus (AAV) in testicular tissue and in abnormal semen samples. Hum Reprod 16: 2333–2337, 2001. doi:10.1093/humrep/16.11.2333.
Crossref | PubMed | Web of Science | Google Scholar - 172. . Bats, primates, and the evolutionary origins and diversification of mammalian gammaherpesviruses. MBio 7:
e01425-16 , 2016. doi:10.1128/mBio.01425-16.
Crossref | PubMed | Web of Science | Google Scholar - 173. . Detection of PCV-2 DNA in stool samples from infants vaccinated with RotaTeq. Hum Vaccin Immunother 10: 25–32, 2014. doi:10.4161/hv.26731.
Crossref | PubMed | Web of Science | Google Scholar - 174. . Does HIV exploit the inflammatory milieu of the male genital tract for successful infection? Front Immunol 7: 245, 2016. doi:10.3389/fimmu.2016.00245.
Crossref | PubMed | Web of Science | Google Scholar - 175. . Characteristics of human sperm chromatin structure following an episode of influenza and high fever: a case study. J Androl 21: 739–746, 2000.
PubMed | Google Scholar - 176. . HIV-RNA decay in paired blood and semen samples of subjects receiving their first dolutegravir-based ART regimen. J Clin Virol 109: 45–49, 2018. doi:10.1016/j.jcv.2018.11.002.
Crossref | PubMed | Web of Science | Google Scholar - 177. . Persistence of free HBV DNA in body secretions and liver despite loss of serum HBV DNA after interferon-induced seroconversion. J Med Virol 20: 183–188, 1986. doi:10.1002/jmv.1890200210.
Crossref | PubMed | Web of Science | Google Scholar - 178. . Comparative observations on intertubular lymphatics and the organization of the interstitial tissue of the mammalian testis. Biol Reprod 9: 500–532, 1973. doi:10.1093/biolreprod/9.5.500.
Crossref | PubMed | Web of Science | Google Scholar - 179. . High resolution in situ hybridization to determine the cellular distribution of lymphocytic choriomeningitis virus RNA in the tissues of persistently infected mice: relevance to arenavirus disease and mechanisms of viral persistence. J Gen Virol 72: 1611–1625, 1991. doi:10.1099/0022-1317-72-7-1611.
Crossref | PubMed | Web of Science | Google Scholar - 180. . Can hepatitis B virus DNA in semen be predicted by serum levels of hepatitis B virus DNA, HBeAg, and HBsAg in chronically infected men from infertile couples? Andrology 3: 506–511, 2015. doi:10.1111/andr.12021.
Crossref | PubMed | Web of Science | Google Scholar - 181. . Timing of intermittent seminal HIV-1 RNA shedding in patients with undetectable plasma viral load under combination antiretroviral therapy. PLoS One 9:
e88922 , 2014. doi:10.1371/journal.pone.0088922.
Crossref | PubMed | Web of Science | Google Scholar - 182. . A distant trophoblast-specific enhancer controls HLA-G expression at the maternal-fetal interface. Proc Natl Acad Sci USA 113: 5364–5369, 2016. doi:10.1073/pnas.1602886113.
Crossref | PubMed | Web of Science | Google Scholar - 183. . Endogenous viruses: insights into viral evolution and impact on host biology. Nat Rev Genet 13: 283–296, 2012. doi:10.1038/nrg3199.
Crossref | PubMed | Web of Science | Google Scholar - 184. . Influence of testosterone on inflammatory response in testicular cells and expression of transcription factor Foxp3 in T cells. Am J Reprod Immunol 74: 12–25, 2015. doi:10.1111/aji.12363.
Crossref | PubMed | Web of Science | Google Scholar - 185. . Infectious, inflammatory and ‘autoimmune’ male factor infertility: how do rodent models inform clinical practice? Hum Reprod Update 24: 416–441, 2018. doi:10.1093/humupd/dmy009.
Crossref | PubMed | Web of Science | Google Scholar - 186. . Testosterone replacement effectively inhibits the development of experimental autoimmune orchitis in rats: evidence for a direct role of testosterone on regulatory T cell expansion. J Immunol 186: 5162–5172, 2011. doi:10.4049/jimmunol.1001958.
Crossref | PubMed | Web of Science | Google Scholar - 187. . Ebola virus ribonucleic acid detection in semen more than two years after resolution of acute Ebola virus infection. Open Forum Infect Dis 4:
ofx155 , 2017. doi:10.1093/ofid/ofx155.
Crossref | PubMed | Web of Science | Google Scholar - 188. . HIV-1 infection of human penile explant tissue and protection by candidate microbicides. AIDS 23: 319–328, 2009. doi:10.1097/QAD.0b013e328321b778.
Crossref | PubMed | Web of Science | Google Scholar - 189. . Duration of Zika viremia in serum. Clin Infect Dis 67: 1143–1144, 2018. doi:10.1093/cid/ciy261.
Crossref | PubMed | Web of Science | Google Scholar - 190. . HPV prophylactic vaccination in males improves the clearance of semen infection. EBioMedicine 2: 1487–1493, 2015. doi:10.1016/j.ebiom.2015.09.005.
Crossref | PubMed | Web of Science | Google Scholar - 191. . HPV-DNA sperm infection and infertility: from a systematic literature review to a possible clinical management proposal. Andrology 3: 163–173, 2015. doi:10.1111/andr.284.
Crossref | PubMed | Web of Science | Google Scholar - 192. . Mechanism of human papillomavirus binding to human spermatozoa and fertilizing ability of infected spermatozoa. PLoS One 6:
e15036 , 2011. doi:10.1371/journal.pone.0015036.
Crossref | PubMed | Web of Science | Google Scholar - 193. . Molecular identification of lyso-glycerophosphocholines as endogenous immunosuppressives in bovine and rat gonadal fluids. Biol Reprod 79: 525–536, 2008. doi:10.1095/biolreprod.107.064386.
Crossref | PubMed | Web of Science | Google Scholar - 194. . Impaired sperm motility in HIV-infected men: an unexpected adverse effect of efavirenz? Hum Reprod 30: 1797–1806, 2015. doi:10.1093/humrep/dev141.
Crossref | PubMed | Web of Science | Google Scholar - 195. . Long-term kinetics of Zika virus RNA and antibodies in body fluids of a vasectomized traveller returning from Martinique: a case report. BMC Infect Dis 17: 55, 2017. doi:10.1186/s12879-016-2123-9.
Crossref | PubMed | Web of Science | Google Scholar - 196. . In vitro models for deciphering the mechanisms underlying the sexual transmission of viruses at the mucosal level. Virology 515: 1–10, 2018. doi:10.1016/j.virol.2017.11.023.
Crossref | PubMed | Web of Science | Google Scholar - 197. . Inhibition of Borna disease virus replication by an endogenous bornavirus-like element in the ground squirrel genome. Proc Natl Acad Sci USA 111: 13175–13180, 2014. doi:10.1073/pnas.1407046111.
Crossref | PubMed | Web of Science | Google Scholar - 198. . Cytokine networks dysregulation during HTLV-1 infection and associated diseases. Viruses 10: 691, 2018. doi:10.3390/v10120691.
Crossref | PubMed | Web of Science | Google Scholar - 199. . Evidence of porcine epidemic diarrhea virus (PEDV) shedding in semen from infected specific pathogen-free boars. Vet Res (Faisalabad) 49: 7, 2018. doi:10.1186/s13567-018-0505-2.
Crossref | PubMed | Web of Science | Google Scholar - 200. . HIV-1 reservoirs in urethral macrophages of patients under suppressive antiretroviral therapy. Nat Microbiol 4: 633–644, 2019. doi:10.1038/s41564-018-0335-z.
Crossref | PubMed | Web of Science | Google Scholar - 201. . The adult penile urethra is a novel entry site for HIV-1 that preferentially targets resident urethral macrophages. Mucosal Immunol 6: 776–786, 2013. doi:10.1038/mi.2012.116.
Crossref | PubMed | Web of Science | Google Scholar - 202. . Within 1 h, HIV-1 uses viral synapses to enter efficiently the inner, but not outer, foreskin mucosa and engages Langerhans-T cell conjugates. Mucosal Immunol 3: 506–522, 2010. doi:10.1038/mi.2010.32.
Crossref | PubMed | Web of Science | Google Scholar - 203. . HIV-1-RNA in seminal plasma correlates with detection of HIV-1-DNA in semen cells, but not with CMV shedding, among MSM on successful antiretroviral regimens. J Antimicrob Chemother 71: 3202–3205, 2016. doi:10.1093/jac/dkw271.
Crossref | PubMed | Web of Science | Google Scholar - 204. . Prevention and control of Zika as a mosquito-borne and sexually transmitted disease: a mathematical modeling analysis. Sci Rep 6: 28070, 2016. doi:10.1038/srep28070.
Crossref | PubMed | Web of Science | Google Scholar - 205. . Murine Sertoli cells promote the development of tolerogenic dendritic cells: a pivotal role of galectin-1. Immunology 148: 253–265, 2016. doi:10.1111/imm.12598.
Crossref | PubMed | Web of Science | Google Scholar - 206. . Development of a repeat-exposure penile SHIV infection model in macaques to evaluate biomedical preventions against HIV. PLoS One 13:
e0194837 , 2018. doi:10.1371/journal.pone.0194837.
Crossref | PubMed | Web of Science | Google Scholar - 207. . Persistence and infectivity of Zika virus in semen after returning from endemic areas: Report of 5 cases. J Clin Virol 96: 110–115, 2017. doi:10.1016/j.jcv.2017.10.006.
Crossref | PubMed | Web of Science | Google Scholar - 208. . Vertical transmission of Border disease infection. J Comp Pathol 91: 467–470, 1981. doi:10.1016/0021-9975(81)90020-7.
Crossref | PubMed | Web of Science | Google Scholar - 209. . Spontaneous fertility and in vitro fertilization outcome: new evidence of human papillomavirus sperm infection. Fertil Steril 105: 65–72.e1, 2016. doi:10.1016/j.fertnstert.2015.09.018.
Crossref | PubMed | Web of Science | Google Scholar - 210. . Sperm viral infection and male infertility: focus on HBV, HCV, HIV, HPV, HSV, HCMV, and AAV. J Reprod Immunol 100: 20–29, 2013. doi:10.1016/j.jri.2013.03.004.
Crossref | PubMed | Web of Science | Google Scholar - 211. . Association, prevalence, and clearance of human papillomavirus and antisperm antibodies in infected semen samples from infertile patients. Fertil Steril 99: 125–131.e2, 2013. doi:10.1016/j.fertnstert.2012.09.006.
Crossref | PubMed | Web of Science | Google Scholar - 212. . Human Papillomavirus Prophylactic Vaccination improves reproductive outcome in infertile patients with HPV semen infection: a retrospective study. Sci Rep 8: 912, 2018. doi:10.1038/s41598-018-19369-z.
Crossref | PubMed | Web of Science | Google Scholar - 213. . Semen characteristics in human immunodeficiency virus (HIV)- and hepatitis C (HCV)-seropositive males: predictors of the success of viral removal after sperm washing. Hum Reprod 20: 1028–1034, 2005. doi:10.1093/humrep/deh699.
Crossref | PubMed | Web of Science | Google Scholar - 214. . Expression of Toll-like receptor 4 in the prostate gland and its association with the severity of prostate cancer. Prostate 69: 1387–1397, 2009. doi:10.1002/pros.20984.
Crossref | PubMed | Web of Science | Google Scholar - 215. . Prostate epithelial cells can act as early sensors of infection by up-regulating TLR4 expression and proinflammatory mediators upon LPS stimulation. J Leukoc Biol 79: 989–998, 2006. doi:10.1189/jlb.1005597.
Crossref | PubMed | Web of Science | Google Scholar - 216. . A TaqMan-based real-time PCR for detection and quantification of porcine parvovirus 4. J Virol Methods 219: 14–17, 2015. doi:10.1016/j.jviromet.2015.03.011.
Crossref | PubMed | Web of Science | Google Scholar - 217. . Peripheral tissue surveillance and residency by memory T cells. Trends Immunol 34: 27–32, 2013. doi:10.1016/j.it.2012.08.008.
Crossref | PubMed | Web of Science | Google Scholar - 218. . Vertical transmission of hepatitis B virus: challenges and solutions. Int J Womens Health 6: 605–611, 2014.
PubMed | Google Scholar - 219. . Central nervous system structures labelled from the testis using the transsynaptic viral tracing technique. J Neuroendocrinol 12: 1087–1095, 2000. doi:10.1046/j.1365-2826.2000.00560.x.
Crossref | PubMed | Web of Science | Google Scholar - 220. . Identification of CNS neurons involved in the innervation of the epididymis: a viral transneuronal tracing study. Auton Neurosci 92: 1–10, 2001. doi:10.1016/S1566-0702(01)00292-2.
Crossref | PubMed | Web of Science | Google Scholar - 221. . Identification of neurones of the brain and spinal cord involved in the innervation of the ductus deferens using the viral tracing method. Int J Androl 26: 91–100, 2003. doi:10.1046/j.1365-2605.2003.00392.x.
Crossref | PubMed | Google Scholar - 222. . Epidemiological aspects and world distribution of HTLV-1 infection. Front Microbiol 3: 388, 2012. doi:10.3389/fmicb.2012.00388.
Crossref | PubMed | Web of Science | Google Scholar - 223. . HIV-1 DNA levels in peripheral blood mononuclear cells and cannabis use are associated with intermittent HIV shedding in semen of men who have sex with men on successful antiretroviral regimens. Clin Infect Dis 58: 1763–1770, 2014. doi:10.1093/cid/ciu187.
Crossref | PubMed | Web of Science | Google Scholar - 224. . Cytomegalovirus DNA in semen and blood is associated with higher levels of proviral HIV DNA. J Infect Dis 207: 898–902, 2013. doi:10.1093/infdis/jis777.
Crossref | PubMed | Web of Science | Google Scholar - 225. . Cytomegalovirus replication in semen is associated with higher levels of proviral HIV DNA and CD4+ T cell activation during antiretroviral treatment. J Virol 88: 7818–7827, 2014. doi:10.1128/JVI.00831-14.
Crossref | PubMed | Web of Science | Google Scholar - 226. . Impact of seminal cytomegalovirus replication on HIV-1 dynamics between blood and semen. J Med Virol 84: 1703–1709, 2012. doi:10.1002/jmv.23398.
Crossref | PubMed | Web of Science | Google Scholar - 227. . Herpes viruses and HIV-1 drug resistance mutations influence the virologic and immunologic milieu of the male genital tract. AIDS 27: 39–47, 2013. doi:10.1097/QAD.0b013e3283573305.
Crossref | PubMed | Web of Science | Google Scholar - 228. . Role of seminal shedding of herpesviruses in HIV type 1 transmission. J Infect Dis 207: 257–261, 2013. doi:10.1093/infdis/jis683.
Crossref | PubMed | Web of Science | Google Scholar - 229. . Shedding of HIV and human herpesviruses in the semen of effectively treated HIV-1-infected men who have sex with men. Clin Infect Dis 57: 441–447, 2013. doi:10.1093/cid/cit252.
Crossref | PubMed | Web of Science | Google Scholar - 230. . Associations between virologic and immunologic dynamics in blood and in the male genital tract. J Virol 86: 1307–1315, 2012. doi:10.1128/JVI.06077-11.
Crossref | PubMed | Web of Science | Google Scholar - 231. . A transitional endogenous lentivirus from the genome of a basal primate and implications for lentivirus evolution. Proc Natl Acad Sci USA 105: 20362–20367, 2008. doi:10.1073/pnas.0807873105.
Crossref | PubMed | Web of Science | Google Scholar - 232. . Parallel germline infiltration of a lentivirus in two Malagasy lemurs. PLoS Genet 5:
e1000425 , 2009. doi:10.1371/journal.pgen.1000425.
Crossref | PubMed | Web of Science | Google Scholar - 233. . Male infertility: a public health issue caused by sexually transmitted pathogens. Nat Rev Urol 11: 672–687, 2014. doi:10.1038/nrurol.2014.285.
Crossref | PubMed | Web of Science | Google Scholar - 234. . Reverse transcriptase activity in mature spermatozoa of mouse. J Cell Biol 148: 1107–1114, 2000. doi:10.1083/jcb.148.6.1107.
Crossref | PubMed | Web of Science | Google Scholar - 235. . Risk factors for HTLV-I and II in individuals attending a clinic for sexually transmitted diseases. Sex Transm Dis 27: 87–92, 2000. doi:10.1097/00007435-200002000-00006.
Crossref | PubMed | Web of Science | Google Scholar - 236. . Age-specific prevalence, incidence, and duration of human papillomavirus infections in a cohort of 290 US men. J Infect Dis 198: 827–835, 2008. doi:10.1086/591095.
Crossref | PubMed | Web of Science | Google Scholar - 237. . Review: Risks of disease transmission through semen in cattle. Animal 12, s1: s165–s171, 2018. doi:10.1017/S1751731118000708.
Crossref | PubMed | Web of Science | Google Scholar - 238. . Pathogens that cause infertility of bulls or transmission via semen. Theriogenology 70: 504–507, 2008. doi:10.1016/j.theriogenology.2008.05.033.
Crossref | PubMed | Web of Science | Google Scholar - 239. . High risk human papilloma viruses (HPVs) are present in benign prostate tissues before development of HPV associated prostate cancer. Infect Agent Cancer 12: 46, 2017. doi:10.1186/s13027-017-0157-2.
Crossref | PubMed | Web of Science | Google Scholar - 240. . Global, regional, and national cancer incidence, mortality, years of life lost, years lived with disability, and disability-adjusted life-years for 29 cancer groups, 1990 to 2016: a systematic analysis for the global burden of disease study. JAMA Oncol 4: 1553–1568, 2018. doi:10.1001/jamaoncol.2018.2706.
Crossref | PubMed | Web of Science | Google Scholar - 241. . Selective changes of retroelement expression in human prostate cancer. Carcinogenesis 32: 1484–1492, 2011. doi:10.1093/carcin/bgr181.
Crossref | PubMed | Web of Science | Google Scholar - 242. . Human endogenous retrovirus HERV-K(HML-2) activity in prostate cancer is dominated by a few loci. Prostate 75: 1958–1971, 2015. doi:10.1002/pros.23095.
Crossref | PubMed | Web of Science | Google Scholar - 243. . Production of the chemokines monocyte chemotactic protein-1, regulated on activation normal T cell expressed and secreted protein, growth-related oncogene, and interferon-gamma-inducible protein-10 is induced by the Sendai virus in human and rat testicular cells. Endocrinology 143: 1434–1440, 2002. doi:10.1210/endo.143.4.8735.
Crossref | PubMed | Web of Science | Google Scholar - 244. . Mumps virus decreases testosterone production and gamma interferon-induced protein 10 secretion by human leydig cells. J Virol 77: 3297–3300, 2003. doi:10.1128/JVI.77.5.3297-3300.2003.
Crossref | PubMed | Web of Science | Google Scholar - 245. . Mucosa: key interactions determining sexual transmission of the HIV infection. Front Immunol 10: 144, 2019. doi:10.3389/fimmu.2019.00144.
Crossref | PubMed | Web of Science | Google Scholar - 246. . Human T cell lymphotropic virus type I infection among female sex workers in Peru. J Infect Dis 169: 754–759, 1994. doi:10.1093/infdis/169.4.754.
Crossref | PubMed | Web of Science | Google Scholar - 247. . Zika virus infection damages the testes in mice. Nature 540: 438–442, 2016. doi:10.1038/nature20556.
Crossref | PubMed | Web of Science | Google Scholar - 248. . Excretion of porcine parvovirus through the genital tract of boars. Am J Vet Res 51: 359–362, 1990.
PubMed | Web of Science | Google Scholar - 249. . HERV envelope proteins: physiological role and pathogenic potential in cancer and autoimmunity. Front Microbiol 9: 462, 2018. doi:10.3389/fmicb.2018.00462.
Crossref | PubMed | Web of Science | Google Scholar - 250. . Evaluation of the transmission of porcine circovirus type 2 (PCV-2) genogroups a and b with semen from infected specific-pathogen-free boars. Vet Microbiol 162: 381–387, 2013. doi:10.1016/j.vetmic.2012.10.011.
Crossref | PubMed | Web of Science | Google Scholar - 251. . Persistence of parvovirus B19 DNA in testis of patients with testicular germ cell tumours. J Gen Virol 79: 573–579, 1998. doi:10.1099/0022-1317-79-3-573.
Crossref | PubMed | Web of Science | Google Scholar - 252. . Male circumcision for HIV prevention in men in Rakai, Uganda: a randomised trial. Lancet 369: 657–666, 2007. doi:10.1016/S0140-6736(07)60313-4.
Crossref | PubMed | Web of Science | Google Scholar - 253. . Male circumcision decreases acquisition and increases clearance of high-risk human papillomavirus in HIV-negative men: a randomized trial in Rakai, Uganda. J Infect Dis 201: 1455–1462, 2010. doi:10.1086/652184.
Crossref | PubMed | Web of Science | Google Scholar - 254. . Matrix fibronectin increases HIV stability and infectivity. J Immunol 168: 5722–5729, 2002. doi:10.4049/jimmunol.168.11.5722.
Crossref | PubMed | Web of Science | Google Scholar - 255. . Viral load, tissue distribution and histopathological lesions in goats naturally and experimentally infected with the Small Ruminant Lentivirus Genotype E (subtype E1 Roccaverano strain). Res Vet Sci 118: 107–114, 2018. doi:10.1016/j.rvsc.2018.01.008.
Crossref | PubMed | Web of Science | Google Scholar - 256. . Detection of type 2 herpes simplex virus in cells of spermatogenic epithelium in infected testes of guinea pigs. Bull Exp Biol Med 144: 73–76, 2007. doi:10.1007/s10517-007-0258-9.
Crossref | PubMed | Web of Science | Google Scholar - 257. . DNA vaccination protects mice against Zika virus-induced damage to the testes. Nat Commun 8: 15743, 2017. doi:10.1038/ncomms15743.
Crossref | PubMed | Web of Science | Google Scholar - 258. . Endogenous retroviruses in the human genome sequence. Genome Biol 2:
1017.1 , 2001. doi:10.1186/gb-2001-2-6-reviews1017.
Crossref | PubMed | Web of Science | Google Scholar - 259. . Demonstration of Epstein-Barr virus in carcinomas of various sites. Cancer Res 62: 4876–4878, 2002.
PubMed | Web of Science | Google Scholar - 260. . What goes around, comes around - HSV-1 replication in monocyte-derived dendritic cells. Front Microbiol 8: 2149, 2017. doi:10.3389/fmicb.2017.02149.
Crossref | PubMed | Web of Science | Google Scholar - 262. . Intrinsic retroviral reactivation in human preimplantation embryos and pluripotent cells. Nature 522: 221–225, 2015. doi:10.1038/nature14308.
Crossref | PubMed | Web of Science | Google Scholar - 263. . Exploring the role of antigen presenting cells in male genital tract. Andrologia 50:
e13120 , 2018. doi:10.1111/and.13120.
Crossref | PubMed | Web of Science | Google Scholar - 264. . The immunobiology of the mammalian epididymis: the black box is now open! Basic Clin Androl 23: 8, 2013. doi:10.1186/2051-4190-23-8.
Crossref | PubMed | Google Scholar - 265. . Human immunodeficiency virus type 1 shedding pattern in semen correlates with the compartmentalization of viral Quasi species between blood and semen. J Infect Dis 182: 79–87, 2000. doi:10.1086/315644.
Crossref | PubMed | Web of Science | Google Scholar - 266. . The epidemiology of Molluscum contagiosum in HIV-seropositive patients: a unique entity or insignificant finding? Int J STD AIDS 19: 503–506, 2008. doi:10.1258/ijsa.2008.008186.
Crossref | PubMed | Web of Science | Google Scholar - 267. . Viral kinetics in semen with different antiretroviral families in treatment-naive human immunodeficiency virus-infected patients: a randomized trial. Clin Infect Dis 65: 551–556, 2017. doi:10.1093/cid/cix358.
Crossref | PubMed | Web of Science | Google Scholar - 268. . [Disorder of spermatogenesis and viral discharge into semen in boars infected with Japanese encephalitis virus (author’s transl)]. Uirusu 27: 21–26, 1977. doi:10.2222/jsv.27.21.
Crossref | PubMed | Google Scholar - 269. . Presence of HBV DNA in spermatozoa: a possible vertical transmission of HBV via the germ line. J Med Virol 16: 61–66, 1985. doi:10.1002/jmv.1890160109.
Crossref | PubMed | Web of Science | Google Scholar - 270. . Retroviruses and the placenta. Curr Biol 22: R609–R613, 2012. doi:10.1016/j.cub.2012.06.002.
Crossref | PubMed | Google Scholar - 271. . Interleukin-1 inhibits Leydig cell steroidogenesis primarily by decreasing 17 alpha-hydroxylase/C17-20 lyase cytochrome P450 expression. Endocrinology 131: 2165–2172, 1992. doi:10.1210/endo.131.5.1425417.
Crossref | PubMed | Web of Science | Google Scholar - 272. . Semen may harbor HIV despite effective HAART: another piece in the puzzle. PLoS One 5:
e10569 , 2010. doi:10.1371/journal.pone.0010569.
Crossref | PubMed | Web of Science | Google Scholar - 273. . Chromosomal integration of human herpesvirus 6 is the major mode of congenital human herpesvirus 6 infection. Pediatrics 122: 513–520, 2008. doi:10.1542/peds.2007-2838.
Crossref | PubMed | Web of Science | Google Scholar - 274. . The effect of pseudorabies (Aujeszky’s) virus infection on young mature boars and boar fertility. Can J Comp Med 48: 192–197, 1984.
PubMed | Google Scholar - 275. . Host defense proteins of the male reproductive tract. J Androl 23: 585–597, 2002.
Crossref | PubMed | Web of Science | Google Scholar - 276. . Biology of Zika virus infection in human skin cells. J Virol 89: 8880–8896, 2015. doi:10.1128/JVI.00354-15.
Crossref | PubMed | Google Scholar - 277. . Rift Valley fever in kidney transplant recipient returning from Mali with viral RNA detected in semen up to four months from symptom onset, France, autumn 2015. Euro Surveill 21: 30222, 2016. doi:10.2807/1560-7917.ES.2016.21.18.30222.
Crossref | PubMed | Google Scholar - 278. . HERV-K(HML-2), a seemingly silent subtenant - but still waters run deep. APMIS 124: 67–87, 2016. doi:10.1111/apm.12475.
Crossref | PubMed | Web of Science | Google Scholar - 279. . Evaluation of the presence of equine viral herpesvirus 1 (EHV-1) and equine viral herpesvirus 4 (EHV-4) DNA in stallion semen using polymerase chain reaction (PCR). Theriogenology 71: 1381–1389, 2009. doi:10.1016/j.theriogenology.2009.01.009.
Crossref | PubMed | Web of Science | Google Scholar - 280. . Prevalence of genital human papillomavirus among men in Europe: systematic review and meta-analysis. J Sex Med 11: 2630–2644, 2014. doi:10.1111/jsm.12652.
Crossref | PubMed | Web of Science | Google Scholar - 281. . Immunophysiology and pathology of inflammation in the testis and epididymis. J Androl 32: 625–640, 2011. doi:10.2164/jandrol.111.012989.
Crossref | PubMed | Web of Science | Google Scholar - 282. . Regulation of activin and inhibin in the adult testis and the evidence for functional roles in spermatogenesis and immunoregulation. Mol Cell Endocrinol 359: 30–42, 2012. doi:10.1016/j.mce.2011.09.031.
Crossref | PubMed | Web of Science | Google Scholar - 283. . Defense and infection of the human placenta. APMIS 126: 570–588, 2018. doi:10.1111/apm.12847.
Crossref | PubMed | Web of Science | Google Scholar - 284. . The role of human papilloma virus in urological malignancies. Anticancer Res 35: 2513–2519, 2015.
PubMed | Web of Science | Google Scholar - 285. . Human herpesvirus 8 establishes a latent infection in prostates of Tobago men resulting in increased macrophage infiltration. Prostate 76: 735–743, 2016. doi:10.1002/pros.23163.
Crossref | PubMed | Web of Science | Google Scholar - 286. . Reduced clearance of penile human papillomavirus infection in uncircumcised men. J Infect Dis 201: 1340–1343, 2010. doi:10.1086/651607.
Crossref | PubMed | Web of Science | Google Scholar - 287. . Zika Virus infection of rhesus macaques leads to viral persistence in multiple tissues. [Correction in PLoS Pathog 13: e1006317, 2017.] PLoS Pathog 13:
e1006219 , 2017. doi:10.1371/journal.ppat.1006219.
Crossref | PubMed | Web of Science | Google Scholar - 288. . Seminal fluid ribavirin level and functional semen parameters in patients with chronic hepatitis C on antiviral combination therapy. J Hepatol 52: 812–816, 2010. doi:10.1016/j.jhep.2009.12.039.
Crossref | PubMed | Web of Science | Google Scholar - 289. . First detection of Schmallenberg virus RNA in bovine semen, Germany, 2012. [Correction in Vet Microbiol 170: 179, 2014.] Vet Microbiol 167: 289–295, 2013. doi:10.1016/j.vetmic.2013.09.002.
Crossref | PubMed | Web of Science | Google Scholar - 289a. . Semen and hormonal parameters in men with chronic hepatitis C infection. Fertil Steril 95: 2557–2559, 2011. doi:10.1016/j.fertnstert.2011.05.014.
Crossref | PubMed | Web of Science | Google Scholar - 290. . Singlet oxygen effects on lipid membranes: implications for the mechanism of action of broad-spectrum viral fusion inhibitors. Biochem J 459: 161–170, 2014. doi:10.1042/BJ20131058.
Crossref | PubMed | Web of Science | Google Scholar - 291. . Defensins in viral infection and pathogenesis. Annu Rev Virol 4: 369–391, 2017. doi:10.1146/annurev-virology-101416-041734.
Crossref | PubMed | Web of Science | Google Scholar - 292. . The evolution of endogenous viral elements. Cell Host Microbe 10: 368–377, 2011. doi:10.1016/j.chom.2011.09.002.
Crossref | PubMed | Web of Science | Google Scholar - 293. . Pathological changes associated with equine arteritis virus infection of the reproductive tract in prepubertal and peripubertal colts. J Comp Pathol 109: 281–293, 1993. doi:10.1016/S0021-9975(08)80253-8.
Crossref | PubMed | Web of Science | Google Scholar - 294. . Histopathologic findings following experimental equine herpesvirus 1 infection of horses. Front Vet Sci 6: 59, 2019. doi:10.3389/fvets.2019.00059.
Crossref | PubMed | Web of Science | Google Scholar - 295. . Endogenous non-retroviral RNA virus elements evidence a novel type of antiviral immunity. Mob Genet Elements 6:
e1165785 , 2016. doi:10.1080/2159256X.2016.1165785.
Crossref | PubMed | Google Scholar - 296. . Trained immunity in newborn infants of HBV-infected mothers. Nat Commun 6: 6588, 2015. doi:10.1038/ncomms7588.
Crossref | PubMed | Web of Science | Google Scholar - 297. . Endogenous non-retroviral RNA virus elements in mammalian genomes. Nature 463: 84–87, 2010. doi:10.1038/nature08695.
Crossref | PubMed | Web of Science | Google Scholar - 298. . Origins of HIV-infected leukocytes and virions in semen. J Infect Dis 210, Suppl 3: S622–S630, 2014. doi:10.1093/infdis/jiu328.
Crossref | PubMed | Web of Science | Google Scholar - 299. . Seminal simian immunodeficiency virus in chronically infected Cynomolgus Macaques is dominated by virus originating from multiple genital organs. J Virol 92:
e00133-18 , 2018. doi:10.1128/JVI.00133-18.
Crossref | PubMed | Web of Science | Google Scholar - 300. . Emerging virus diseases: can we ever expect the unexpected? Emerg Microbes Infect 1:
e46 , 2012. doi:10.1038/emi.2012.47.
Crossref | PubMed | Web of Science | Google Scholar - 301. . Thirty years of research on infection and prostate cancer: no conclusive evidence for a link. A systematic review. Urol Oncol 31: 951–965, 2013. doi:10.1016/j.urolonc.2012.01.013.
Crossref | PubMed | Web of Science | Google Scholar - 302. . Activation of Toll like receptor 3 induces spermatogonial stem cell apoptosis. Cell Biochem Funct 33: 415–422, 2015. doi:10.1002/cbf.3133.
Crossref | PubMed | Web of Science | Google Scholar - 303. . High prevalence of hepatitis E virus in semen of infertile male and causes testis damage. Gut 67: 1199–1201, 2018. doi:10.1136/gutjnl-2017-314884.
Crossref | PubMed | Web of Science | Google Scholar - 304. . Effects of hepatitis B virus infection on human sperm chromosomes. World J Gastroenterol 9: 736–740, 2003. doi:10.3748/wjg.v9.i4.736.
Crossref | PubMed | Web of Science | Google Scholar - 305. . Hepatitis B virus s protein enhances sperm apoptosis and reduces sperm fertilizing capacity in vitro. PLoS One 8:
e68688 , 2013. doi:10.1371/journal.pone.0068688.
Crossref | PubMed | Web of Science | Google Scholar - 306. . Sexually transmissible infections and prostate cancer risk. Cancer Epidemiol Biomarkers Prev 17: 2374–2381, 2008. doi:10.1158/1055-9965.EPI-08-0173.
Crossref | PubMed | Web of Science | Google Scholar - 307. . Zika virus in semen: a prospective cohort study of symptomatic travellers returning to Belgium. Bull World Health Organ 95: 802–809, 2017. doi:10.2471/BLT.17.181370.
Crossref | PubMed | Web of Science | Google Scholar - 308. . Human endogenous retrovirus envelope proteins target dendritic cells to suppress T-cell activation. Eur J Immunol 45: 1748–1759, 2015. doi:10.1002/eji.201445366.
Crossref | PubMed | Web of Science | Google Scholar - 309. . [Mollusca contagiosa in HIV infection. Clinical manifestation, relation to immune status and prognostic value in 39 patients]. Hautarzt 48: 103–109, 1997. doi:10.1007/s001050050554.
Crossref | PubMed | Web of Science | Google Scholar - 310. . Hepatitis B: the immaculate infection. Gastroenterol Hepatol (N Y) 3: 724–726, 2007.
PubMed | Google Scholar - 311. . Human T-cell lymphotropic virus type 1. In: Biological Agents. Lyon: International Agency for Research on Cancer, 2012.
Google Scholar - 312. . Toll like receptor-9 agonists stimulate prostate cancer invasion in vitro. Prostate 67: 774–781, 2007. doi:10.1002/pros.20562.
Crossref | PubMed | Web of Science | Google Scholar - 313. . HIV-1-RNA Decay and Dolutegravir Concentrations in Semen of Patients Starting a First Antiretroviral Regimen. J Infect Dis 214: 1512–1519, 2016. doi:10.1093/infdis/jiw406.
Crossref | PubMed | Web of Science | Google Scholar - 314. . Hepatitis B virus and its sexually transmitted infection - an update. Microb Cell 3: 419–437, 2016. doi:10.15698/mic2016.09.527.
Crossref | PubMed | Web of Science | Google Scholar - 315. . Seminal plasma induces inflammation and enhances HIV-1 replication in human cervical tissue explants. PLoS Pathog 13:
e1006402 , 2017. doi:10.1371/journal.ppat.1006402.
Crossref | PubMed | Web of Science | Google Scholar - 316. . Excretion of lumpy skin disease virus in bull semen. Theriogenology 63: 1290–1297, 2005. doi:10.1016/j.theriogenology.2004.06.013.
Crossref | PubMed | Web of Science | Google Scholar - 317. . Major differences between the testis and epididymis in the induction of granulomas in response to extravasated germ cells. I. A light microscopical study in mice. Int J Androl 22: 316–323, 1999. doi:10.1046/j.1365-2605.1999.00186.x.
Crossref | PubMed | Google Scholar - 318. . Transmission of HTLV-I to rabbits via semen and breast milk from seropositive healthy persons. Int J Cancer 45: 980–983, 1990. doi:10.1002/ijc.2910450534.
Crossref | PubMed | Web of Science | Google Scholar - 319. . Mumps epididymo-orchitis with prolonged detection of virus in semen and the development of anti-sperm antibodies. J Med Virol 73: 147–150, 2004. doi:10.1002/jmv.10544.
Crossref | PubMed | Web of Science | Google Scholar - 320. . Male circumcision and HIV transmission; What do we know? Open AIDS J 8: 31–44, 2014. doi:10.2174/1874613601408010031.
Crossref | PubMed | Google Scholar - 321. . The rat epididymal transcriptome: comparison of segmental gene expression in the rat and mouse epididymides. Biol Reprod 76: 561–570, 2007. doi:10.1095/biolreprod.106.057323.
Crossref | PubMed | Web of Science | Google Scholar - 322. . Immune tolerance properties of the testicular tissue as a viral sanctuary site in ART-treated HIV-infected adults. AIDS 30: 2777–2786, 2016. doi:10.1097/QAD.0000000000001282.
Crossref | PubMed | Web of Science | Google Scholar - 323. . Effects of retroviruses on host genome function. Annu Rev Genet 42: 709–732, 2008. doi:10.1146/annurev.genet.42.110807.091501.
Crossref | PubMed | Web of Science | Google Scholar - 324. . Factors associated with seminal impairment in HIV-infected men under antiretroviral therapy. Hum Reprod 32: 265–271, 2017. doi:10.1093/humrep/dew321.
Crossref | PubMed | Web of Science | Google Scholar - 325. . C-X-C motif chemokine ligand 10 produced by mouse Sertoli cells in response to mumps virus infection induces male germ cell apoptosis. [Correction in Cell Death Dis 9: 789, 2018.] Cell Death Dis 8:
e3146 , 2017. doi:10.1038/cddis.2017.560.
Crossref | PubMed | Web of Science | Google Scholar - 326. . Effect of acute Zika virus infection on sperm and virus clearance in body fluids: a prospective observational study. Lancet Infect Dis 17: 1200–1208, 2017. doi:10.1016/S1473-3099(17)30444-9.
Crossref | PubMed | Web of Science | Google Scholar - 327. . Cell-free HTLV-1 infects dendritic cells leading to transmission and transformation of CD4(+) T cells. Nat Med 14: 429–436, 2008. doi:10.1038/nm1745.
Crossref | PubMed | Web of Science | Google Scholar - 328. . Seminal plasma induces prostaglandin-endoperoxide synthase (PTGS) 2 expression in immortalized human vaginal cells: involvement of semen prostaglandin E2 in PTGS2 upregulation. Biol Reprod 88: 13, 2013. doi:10.1095/biolreprod.112.101956.
Crossref | PubMed | Web of Science | Google Scholar - 329. . Human immunodeficiency virus viral load in blood plasma and semen: review and implications of empirical findings. Sex Transm Dis 35: 55–60, 2008. doi:10.1097/OLQ.0b013e318141fe9b.
Crossref | PubMed | Web of Science | Google Scholar - 330. . Activation of Nod1 and Nod2 induces innate immune responses of prostate epithelial cells. Prostate 72: 1351–1358, 2012. doi:10.1002/pros.22483.
Crossref | PubMed | Web of Science | Google Scholar - 331. . Effects of hepatitis B virus S protein exposure on sperm membrane integrity and functions. PLoS One 7:
e33471 , 2012. doi:10.1371/journal.pone.0033471.
Crossref | PubMed | Web of Science | Google Scholar - 332. . Association between testicular cancer and epididymoorchitis: a population-based case-control study. Sci Rep 6: 23079, 2016. doi:10.1038/srep23079.
Crossref | PubMed | Web of Science | Google Scholar - 333. . Discovery and characterization of mammalian endogenous parvoviruses. J Virol 84: 12628–12635, 2010. doi:10.1128/JVI.01732-10.
Crossref | PubMed | Web of Science | Google Scholar - 334. . Detection of herpes simplex virus, cytomegalovirus, and Epstein-Barr virus in the semen of men attending an infertility clinic. Fertil Steril 79, Suppl 3: 1566–1570, 2003. doi:10.1016/S0015-0282(03)00370-4.
Crossref | PubMed | Web of Science | Google Scholar - 335. . No increased sperm DNA fragmentation index in semen containing human papillomavirus or herpesvirus. Andrology 1: 361–364, 2013. doi:10.1111/j.2047-2927.2013.00067.x.
Crossref | PubMed | Web of Science | Google Scholar - 336. . Seminal shedding of human herpesviruses. Virol J 10: 226, 2013. doi:10.1186/1743-422X-10-226.
Crossref | PubMed | Web of Science | Google Scholar - 337. . Endogenous retroviruses and the development of cancer. J Immunol 192: 1343–1349, 2014. doi:10.4049/jimmunol.1302972.
Crossref | PubMed | Web of Science | Google Scholar - 338. . Immune responses to endogenous retroelements: taking the bad with the good. Nat Rev Immunol 16: 207–219, 2016. doi:10.1038/nri.2016.27.
Crossref | PubMed | Web of Science | Google Scholar - 339. . Endogenous viral elements in animal genomes. PLoS Genet 6:
e1001191 , 2010. doi:10.1371/journal.pgen.1001191.
Crossref | PubMed | Web of Science | Google Scholar - 340. . The evolutionary dynamics of endogenous retroviruses. Trends Microbiol 13: 463–468, 2005. doi:10.1016/j.tim.2005.08.004.
Crossref | PubMed | Web of Science | Google Scholar - 341. . Discovery and analysis of the first endogenous lentivirus. Proc Natl Acad Sci USA 104: 6261–6265, 2007. doi:10.1073/pnas.0700471104.
Crossref | PubMed | Web of Science | Google Scholar - 342. . Testisimmune privilege - assumptions versus facts. Anim Reprod 10: 3–15, 2013.
PubMed | Web of Science | Google Scholar - 343. . Sertoli cells--immunological sentinels of spermatogenesis. Semin Cell Dev Biol 30: 36–44, 2014. doi:10.1016/j.semcdb.2014.02.011.
Crossref | PubMed | Web of Science | Google Scholar - 344. . Demonstration of MDR1 P-glycoprotein isoform expression in benign and malignant human prostate cells by isoform-specific monoclonal antibodies. Cancer Lett 150: 147–153, 2000. doi:10.1016/S0304-3835(99)00384-5.
Crossref | PubMed | Web of Science | Google Scholar - 345. . Tissue tropisms, infection kinetics, histologic lesions, and antibody response of the MR766 strain of Zika virus in a murine model. Virol J 14: 82, 2017. doi:10.1186/s12985-017-0749-x.
Crossref | PubMed | Web of Science | Google Scholar - 346. . Identification of a RELIK orthologue in the European hare (Lepus europaeus) reveals a minimum age of 12 million years for the lagomorph lentiviruses. Virology 384: 7–11, 2009. doi:10.1016/j.virol.2008.10.045.
Crossref | PubMed | Web of Science | Google Scholar - 347. . Detection of swine Torque teno virus genogroups 1 and 2 in boar sera and semen. Theriogenology 68: 966–971, 2007. doi:10.1016/j.theriogenology.2007.07.010.
Crossref | PubMed | Web of Science | Google Scholar - 348. . West Nile Virus Meningo-Encephalitis: possible sexual transmission. J La State Med Soc 168: 21–22, 2016.
PubMed | Google Scholar - 349. . Insights into Ebola and other emerging and re-emerging infectious disease risks. Environ Health Insights 8:
S20757 , 2014. doi:10.4137/EHI.S20757.
Crossref | PubMed | Google Scholar - 350. . Persistent bovine viral diarrhea virus (BVDV) infection in cattle herds. Iran J Vet Res 18: 154–163, 2017.
PubMed | Web of Science | Google Scholar - 351. . Review. Gastroenterol Hepatol (N Y) 3: 727–730, 2007.
PubMed | Google Scholar - 352. . Relationship between human papillomavirus and penile cancer-implications for prevention and treatment. Transl Androl Urol 6: 791–802, 2017. doi:10.21037/tau.2017.06.27.
Crossref | PubMed | Web of Science | Google Scholar - 353. . Retroviruses and reproduction revisited. J Assist Reprod Genet 35: 1969–1972, 2018. doi:10.1007/s10815-018-1253-y.
Crossref | PubMed | Web of Science | Google Scholar - 354. . Simultaneous detection and differentiation between porcine circovirus and porcine parvovirus in boar semen by multiplex seminested polymerase chain reaction. J Vet Med Sci 65: 741–744, 2003. doi:10.1292/jvms.65.741.
Crossref | PubMed | Web of Science | Google Scholar - 355. . Semen-mediated enhancement of HIV infection is donor-dependent and correlates with the levels of SEVI. Retrovirology 7: 55, 2010. doi:10.1186/1742-4690-7-55.
Crossref | PubMed | Web of Science | Google Scholar - 356. . [Cytomegaloviral inclusions detected in the seminal vesicle, ductus deferens and lungs in an autopsy case of lung cancer]. Rinsho Byori 41: 1059–1062, 1993.
PubMed | Google Scholar - 357. . Sexual transmission efficiency of hepatitis B virus and human immunodeficiency virus among homosexual men. JAMA 264: 230–234, 1990. doi:10.1001/jama.1990.03450020082031.
Crossref | PubMed | Web of Science | Google Scholar - 358. . Replication of bovine viral diarrhoea virus in the bovine reproductive tract and excretion of virus in semen during acute and chronic infections. Vet Rec 128: 587–590, 1991. doi:10.1136/vr.128.25.587.
Crossref | PubMed | Web of Science | Google Scholar - 359. . Long-term persistence of human papillomavirus in the skin of the glans penis of elderly men above 80 years of age. Int J STD AIDS 29: 552–556, 2018. doi:10.1177/0956462417742319.
Crossref | PubMed | Web of Science | Google Scholar - 360. . The Y chromosome: a graveyard for endogenous retroviruses. Gene 161: 163–170, 1995. doi:10.1016/0378-1119(95)00248-5.
Crossref | PubMed | Web of Science | Google Scholar - 361. . Inactivation of human immunodeficiency virus type 1 by the amine oxidase-peroxidase system. J Clin Microbiol 33: 2054–2057, 1995. doi:10.1128/JCM.33.8.2054-2057.1995.
Crossref | PubMed | Web of Science | Google Scholar - 362. . A non-retroviral RNA virus persists in DNA form. Nature 390: 298–301, 1997. doi:10.1038/36876.
Crossref | PubMed | Web of Science | Google Scholar - 363. . Herpes simplex: insights on pathogenesis and possible vaccines. Annu Rev Med 59: 381–395, 2008. doi:10.1146/annurev.med.59.061606.095540.
Crossref | PubMed | Web of Science | Google Scholar - 364. . Development of a Zika virus infection model in Cynomolgus Macaques. Front Microbiol 7: 2028, 2016. doi:10.3389/fmicb.2016.02028.
Crossref | PubMed | Web of Science | Google Scholar - 365. . Long-term ex vivo maintenance of testis tissues producing fertile sperm in a microfluidic device. Sci Rep 6: 21472, 2016. doi:10.1038/srep21472.
Crossref | PubMed | Web of Science | Google Scholar - 366. . Relationship between the mechanism of hepatitis B virus father-infant transmission and pregnancy outcome. Arch Gynecol Obstet 295: 253–257, 2017. doi:10.1007/s00404-016-4231-6.
Crossref | PubMed | Web of Science | Google Scholar - 367. . DC-SIGN: escape mechanism for pathogens. Nat Rev Immunol 3: 697–709, 2003. doi:10.1038/nri1182.
Crossref | PubMed | Web of Science | Google Scholar - 368. . Human semen or seminal plasma does not enhance HIV-1BaL ex vivo infection of human colonic explants. AIDS Res Hum Retroviruses 34: 459–466, 2018. doi:10.1089/aid.2017.0118.
Crossref | PubMed | Web of Science | Google Scholar - 369. . Site-specific integration by adeno-associated virus. Proc Natl Acad Sci USA 87: 2211–2215, 1990. doi:10.1073/pnas.87.6.2211.
Crossref | PubMed | Google Scholar - 370. . Detection of herpes simplex virus DNA in human spermatozoa by in situ hybridization technique. In Vivo 12: 391–394, 1998.
PubMed | Web of Science | Google Scholar - 371. . Virus infections in the nervous system. Cell Host Microbe 13: 379–393, 2013. doi:10.1016/j.chom.2013.03.010.
Crossref | PubMed | Web of Science | Google Scholar - 372. . Hantavirus cardiopulmonary syndrome due to imported Andes Hantavirus infection in Switzerland: a multidisciplinary challenge, two cases and a literature review. Clin Infect Dis 67: 1788–1795, 2018. doi:10.1093/cid/ciy443.
Crossref | PubMed | Web of Science | Google Scholar - 373. . Effects of herpes simplex virus infections on seminal parameters in male partners of infertile couples. Urology 113: 52–58, 2018. doi:10.1016/j.urology.2017.11.050.
Crossref | PubMed | Web of Science | Google Scholar - 374. . Human endogenous retroviruses in neurological diseases. Trends Mol Med 24: 379–394, 2018. doi:10.1016/j.molmed.2018.02.007.
Crossref | PubMed | Web of Science | Google Scholar - 375. . Determinants of prevalence, acquisition, and persistence of human papillomavirus in healthy Mexican military men. Cancer Epidemiol Biomarkers Prev 14: 1710–1716, 2005. doi:10.1158/1055-9965.EPI-04-0926.
Crossref | PubMed | Web of Science | Google Scholar - 376. . The expression of TLR2 and TLR3 in sertoli cells of azoospermic patients. Cell J 19: 375–385, 2017. doi:10.22074/cellj.2017.4300.
Crossref | PubMed | Web of Science | Google Scholar - 377. . Prolonged detection of dengue virus RNA in the semen of a man returning from Thailand to Italy, January 2018. Euro Surveill 23: 23, 2018. doi:10.2807/1560-7917.ES.2018.23.18.18-00197.
Crossref | PubMed | Google Scholar - 378. . [Distribution of hepatitis B virus in testicle tissue in patients with hepatitis B infection]. Zhonghua Yi Xue Za Zhi 73: 329–331, 1993.
PubMed | Google Scholar - 379. . Prevalence of human papillomaviruses in semen: a systematic review and meta-analysis. Hum Reprod 29: 640–651, 2014. doi:10.1093/humrep/det453.
Crossref | PubMed | Web of Science | Google Scholar - 380. . PCR detection and evidence of shedding of porcine circovirus type 2 in boar semen. J Clin Microbiol 38: 4629–4632, 2000. doi:10.1128/JCM.38.12.4629-4632.2000.
Crossref | PubMed | Web of Science | Google Scholar - 381. . Kinetics of Zika viral load in semen. N Engl J Med 377: 697–699, 2017. doi:10.1056/NEJMc1612600.
Crossref | PubMed | Web of Science | Google Scholar - 382. . Impact of positive hepatitis B surface antigen on the outcome of IVF treatment. Reprod Biomed Online 21: 712–717, 2010. doi:10.1016/j.rbmo.2010.06.036.
Crossref | PubMed | Web of Science | Google Scholar - 383. . Viral RNA load in semen from bluetongue serotype 8-infected rams: relationship with sperm quality. Vet J 192: 304–310, 2012. doi:10.1016/j.tvjl.2011.06.028.
Crossref | PubMed | Web of Science | Google Scholar - 384. . Effects of antiretroviral therapy on semen quality. AIDS 22: 637–642, 2008. doi:10.1097/QAD.0b013e3282f4de10.
Crossref | PubMed | Web of Science | Google Scholar - 385. . The inner foreskin of healthy males at risk of HIV infection harbors epithelial CD4+ CCR5+ cells and has features of an inflamed epidermal barrier. PLoS One 9:
e108954 , 2014. doi:10.1371/journal.pone.0108954.
Crossref | PubMed | Web of Science | Google Scholar - 386. . Detection of hepatitis C virus in the semen of infected men. Lancet 356: 42–43, 2000. doi:10.1016/S0140-6736(00)02435-1.
Crossref | PubMed | Web of Science | Google Scholar - 387. . Is sperm evaluation useful in predicting human fertility? Reproduction 134: 31–40, 2007. doi:10.1530/REP-07-0152.
Crossref | PubMed | Web of Science | Google Scholar - 388. . Prevalence of hepatitis E virus (HEV) infection in various pig farms from Shaanxi Province, China: first detection of HEV RNA in pig semen. Transbound Emerg Dis 66: 72–82, 2019. doi:10.1111/tbed.12966.
Crossref | PubMed | Web of Science | Google Scholar - 389. . Multiple diverse circoviruses infect farm animals and are commonly found in human and chimpanzee feces. J Virol 84: 1674–1682, 2010. doi:10.1128/JVI.02109-09.
Crossref | PubMed | Web of Science | Google Scholar - 390. . Characterization of a 2016 clinical isolate of Zika virus in non-human primates. EBioMedicine 12: 170–177, 2016. doi:10.1016/j.ebiom.2016.09.022.
Crossref | PubMed | Web of Science | Google Scholar - 391. . The first case of sexual transmission of dengue in Spain. J Travel Med 27:
taz087 , 2020. doi:10.1093/jtm/taz087.
Crossref | PubMed | Web of Science | Google Scholar - 392. . Herpes simplex virus-specific CD8+ T cells can clear established lytic infections from skin and nerves and can partially limit the early spread of virus after cutaneous inoculation. J Immunol 172: 392–397, 2004. doi:10.4049/jimmunol.172.1.392.
Crossref | PubMed | Web of Science | Google Scholar - 393. . Natural Inhibitor of human cytomegalovirus in human seminal plasma. J Virol 93:
e01855-18 , 2019. doi:10.1128/JVI.01855-18.
Crossref | PubMed | Web of Science | Google Scholar - 394. . Semen of HIV-1-infected individuals: local shedding of herpesviruses and reprogrammed cytokine network. J Infect Dis 205: 97–105, 2012. doi:10.1093/infdis/jir700.
Crossref | PubMed | Web of Science | Google Scholar - 395. . The semen microbiome and its relationship with local immunology and viral load in HIV infection. PLoS Pathog 10:
e1004262 , 2014. doi:10.1371/journal.ppat.1004262.
Crossref | PubMed | Web of Science | Google Scholar - 396. . Penile Anaerobic Dysbiosis as a Risk Factor for HIV Infection. MBio 8:
e00996-17 , 2017. doi:10.1128/mBio.00996-17.
Crossref | PubMed | Web of Science | Google Scholar - 397. . Widespread endogenization of densoviruses and parvoviruses in animal and human genomes. J Virol 85: 9863–9876, 2011. doi:10.1128/JVI.00828-11.
Crossref | PubMed | Web of Science | Google Scholar - 398. . Viral load in the natural history of human papillomavirus infection among men in rural China: a population-based prospective study. Clin Infect Dis 67: 1861–1867, 2018. doi:10.1093/cid/ciy376.
Crossref | PubMed | Web of Science | Google Scholar - 399. . Homozygous defect in HIV-1 coreceptor accounts for resistance of some multiply-exposed individuals to HIV-1 infection. Cell 86: 367–377, 1996. doi:10.1016/S0092-8674(00)80110-5.
Crossref | PubMed | Web of Science | Google Scholar - 400. . Molecular phylogenetic and positive selection analysis of Japanese encephalitis virus strains isolated from pigs in China. Virus Res 178: 547–552, 2013. doi:10.1016/j.virusres.2013.09.002.
Crossref | PubMed | Web of Science | Google Scholar - 401. . Degradation and remobilization of endogenous retroviruses by recombination during the earliest stages of a germ-line invasion. Proc Natl Acad Sci USA 115: 8609–8614, 2018. doi:10.1073/pnas.1807598115.
Crossref | PubMed | Web of Science | Google Scholar - 402. . Impact of chronic viral diseases on semen parameters. Andrologia 42: 121–126, 2010. doi:10.1111/j.1439-0272.2009.00970.x.
Crossref | PubMed | Web of Science | Google Scholar - 403. . Semen lactoferrin promotes CCL20 production by epithelial cells: involvement in HIV transmission. World J Virol 3: 11–17, 2014. doi:10.5501/wjv.v3.i2.11.
Crossref | PubMed | Google Scholar - 404. . Sertoli cells for cell transplantation: pre-clinical studies and future perspectives. Andrology 6: 385–395, 2018. doi:10.1111/andr.12484.
Crossref | PubMed | Web of Science | Google Scholar - 405. . Neurotropic virus infections as the cause of immediate and delayed neuropathology. Acta Neuropathol 131: 159–184, 2016. doi:10.1007/s00401-015-1511-3.
Crossref | PubMed | Web of Science | Google Scholar - 406. . Effect of sexual transmission on the West Africa Ebola outbreak in 2014: a mathematical modelling study. Sci Rep 9: 1653, 2019. doi:10.1038/s41598-018-38397-3.
Crossref | PubMed | Web of Science | Google Scholar - 407. . Presence of human papillomavirus in semen of healthy men is firmly associated with HPV infections of the penile epithelium. Fertil Steril 104: 838–844.e8, 2015. doi:10.1016/j.fertnstert.2015.06.028.
Crossref | PubMed | Web of Science | Google Scholar - 408. . Presence of human papillomavirus in semen in relation to semen quality. Hum Reprod 31:
dev317 , 2016. doi:10.1093/humrep/dev317.
Crossref | PubMed | Web of Science | Google Scholar - 409. . Human papillomavirus in semen and the risk for male infertility: a systematic review and meta-analysis. BMC Infect Dis 17: 714, 2017. doi:10.1186/s12879-017-2812-z.
Crossref | PubMed | Web of Science | Google Scholar - 410. . Zika virus causes testis damage and leads to male infertility in mice. [Correction in Cell 168: 542, 2017.] Cell 167: 1511–1524.e10, 2016. doi:10.1016/j.cell.2016.11.016.
Crossref | PubMed | Web of Science | Google Scholar - 411. . Lymphatic dissemination of simian immunodeficiency virus after penile inoculation. J Virol 90: 4093–4104, 2016. doi:10.1128/JVI.02947-15.
Crossref | PubMed | Web of Science | Google Scholar - 412. . Genome-wide amplification of proviral sequences reveals new polymorphic HERV-K(HML-2) proviruses in humans and chimpanzees that are absent from genome assemblies. Retrovirology 12: 35, 2015. doi:10.1186/s12977-015-0162-8.
Crossref | PubMed | Web of Science | Google Scholar - 413. . Innate immunity in the male genital tract: Chlamydia trachomatis induces keratinocyte-derived chemokine production in prostate, seminal vesicle and epididymis/vas deferens primary cultures. J Med Microbiol 60: 307–316, 2011. doi:10.1099/jmm.0.024877-0.
Crossref | PubMed | Web of Science | Google Scholar - 414. . Susceptibility of prostate epithelial cells to Chlamydia muridarum infection and their role in innate immunity by recruitment of intracellular Toll-like receptors 4 and 2 and MyD88 to the inclusion. Infect Immun 74: 6973–6981, 2006. doi:10.1128/IAI.00593-06.
Crossref | PubMed | Web of Science | Google Scholar - 415. . Effect of chickenpox and of pneumonia on semen quality. Fertil Steril 2: 523–533, 1951. doi:10.1016/S0015-0282(16)30726-9.
Crossref | PubMed | Web of Science | Google Scholar - 416. . Characterization of shedding patterns of Porcine circovirus types 2a and 2b in experimentally inoculated mature boars. J Vet Diagn Invest 20: 725–734, 2008. doi:10.1177/104063870802000603.
Crossref | PubMed | Web of Science | Google Scholar - 417. . ‘There and back again’: revisiting the pathophysiological roles of human endogenous retroviruses in the post-genomic era. Philos Trans R Soc Lond B Biol Sci 368:
20120504 , 2013. doi:10.1098/rstb.2012.0504.
Crossref | PubMed | Web of Science | Google Scholar - 418. . Sperm/DNA interaction: integration of foreign DNA sequences in the mouse sperm genome. J Reprod Immunol 41: 187–196, 1998. doi:10.1016/S0165-0378(98)00058-8.
Crossref | PubMed | Web of Science | Google Scholar - 419. . HIV binding, penetration, and primary infection in human cervicovaginal tissue. Proc Natl Acad Sci USA 102: 11504–11509, 2005. doi:10.1073/pnas.0500848102.
Crossref | PubMed | Web of Science | Google Scholar - 420. . Activation of endogenous nucleases in mature sperm cells upon interaction with exogenous DNA. DNA Cell Biol 16: 1087–1097, 1997. doi:10.1089/dna.1997.16.1087.
Crossref | PubMed | Web of Science | Google Scholar - 421. . Fighting fire with fire: endogenous retrovirus envelopes as restriction factors. J Virol 89: 4047–4050, 2015. doi:10.1128/JVI.03653-14.
Crossref | PubMed | Web of Science | Google Scholar - 422. . Variation of semen quality in normal men. Int J Androl 14: 99–107, 1991. doi:10.1111/j.1365-2605.1991.tb01071.x.
Crossref | PubMed | Google Scholar - 423. . Herpes simplex virus inoculation in murine rete testis results in irreversible testicular damage. Int J Exp Pathol 95: 120–130, 2014. doi:10.1111/iep.12071.
Crossref | PubMed | Web of Science | Google Scholar - 424. . Characterization of Marburg virus glycoprotein in viral entry. Virology 358: 79–88, 2007. doi:10.1016/j.virol.2006.06.041.
Crossref | PubMed | Web of Science | Google Scholar - 425. . Mumps orchitis. Urology 36: 355–358, 1990. doi:10.1016/0090-4295(90)80248-L.
Crossref | PubMed | Web of Science | Google Scholar - 426. . Peripheral plasma and semen cytokine response to Zika virus in humans. Emerg Infect Dis 25: 823–825, 2019. doi:10.3201/eid2504.171886.
Crossref | PubMed | Web of Science | Google Scholar - 427. . Zika virus in semen and spermatozoa. Lancet Infect Dis 16: 1106–1107, 2016. doi:10.1016/S1473-3099(16)30336-X.
Crossref | PubMed | Web of Science | Google Scholar - 428. . Reappraising the microscopic anatomy of human testis: identification of telocyte networks in the peritubular and intertubular stromal space. Sci Rep 8: 14780, 2018. doi:10.1038/s41598-018-33126-2.
Crossref | PubMed | Web of Science | Google Scholar - 429. . Cationic polypeptides contribute to the anti-HIV-1 activity of human seminal plasma. FASEB J 23: 3609–3618, 2009. doi:10.1096/fj.09-131961.
Crossref | PubMed | Web of Science | Google Scholar - 430. . Tissue and cellular tropism, pathology and pathogenesis of Ebola and Marburg viruses. J Pathol 235: 153–174, 2015. doi:10.1002/path.4456.
Crossref | PubMed | Web of Science | Google Scholar - 431. . SV40 early region and large T antigen in human brain tumors, peripheral blood cells, and sperm fluids from healthy individuals. Cancer Res 56: 4820–4825, 1996.
PubMed | Web of Science | Google Scholar - 432. . Papilloma and polyoma DNA tumor virus sequences in female genital tumors. Cancer Invest 22: 697–705, 2004. doi:10.1081/CNV-200032937.
Crossref | PubMed | Web of Science | Google Scholar - 433. . Marburg virus disease. Postgrad Med J 49: 542–546, 1973. doi:10.1136/pgmj.49.574.542.
Crossref | PubMed | Web of Science | Google Scholar - 434. . [Spermatogenic transmission of the “Marburg virus” (causes of “Marburg simian disease”)]. Klin Wochenschr 46: 398–400, 1968. doi:10.1007/BF01734141.
Crossref | PubMed | Google Scholar - 435. . Presence of human papillomavirus DNA in testicular biopsies from nonobstructive azoospermic men. Arch Pathol Lab Med 129: 1132–1136, 2005.
PubMed | Web of Science | Google Scholar - 436. . Inflammation in prostate carcinogenesis. Nat Rev Cancer 7: 256–269, 2007. doi:10.1038/nrc2090.
Crossref | PubMed | Web of Science | Google Scholar - 437. . Mumps orchitis. J R Soc Med 99: 573–575, 2006. doi:10.1177/014107680609901116.
Crossref | PubMed | Web of Science | Google Scholar - 438. . Hepatitis B virus replication in diverse cell types during chronic hepatitis B virus infection. Hepatology 18: 781–789, 1993. doi:10.1002/hep.1840180406.
Crossref | PubMed | Google Scholar - 439. . Existence of TT virus DNA in extracellular body fluids from normal healthy Japanese subjects. Intervirology 43: 16–19, 2000. doi:10.1159/000025018.
Crossref | PubMed | Web of Science | Google Scholar - 440. . Detection of simian immunodeficiency virus in semen, urethra, and male reproductive organs during efficient highly active antiretroviral therapy. J Virol 89: 5772–5787, 2015. doi:10.1128/JVI.03628-14.
Crossref | PubMed | Web of Science | Google Scholar - 441. . Zika virus infects human testicular tissue and germ cells. J Clin Invest 128: 4697–4710, 2018. doi:10.1172/JCI121735.
Crossref | PubMed | Web of Science | Google Scholar - 442. . Sterile inflammation as a factor in human male infertility: Involvement of Toll like receptor 2, biglycan and peritubular cells. Sci Rep 6: 37128, 2016. doi:10.1038/srep37128.
Crossref | PubMed | Web of Science | Google Scholar - 443. . Resistance of castrated male horses to attempted establishment of the carrier state with equine arteritis virus. J Comp Pathol 111: 383–388, 1994. doi:10.1016/S0021-9975(05)80096-9.
Crossref | PubMed | Web of Science | Google Scholar - 444. . Potential HIV-1 target cells in the human penis. AIDS 20: 1491–1495, 2006. doi:10.1097/01.aids.0000237364.11123.98.
Crossref | PubMed | Web of Science | Google Scholar - 445. . Pathogenesis and sexual transmission of Spondweni and Zika viruses. PLoS Negl Trop Dis 11:
e0005990 , 2017. doi:10.1371/journal.pntd.0005990.
Crossref | PubMed | Web of Science | Google Scholar - 446. . Adolescent and adult risk factors for testicular cancer. Nat Rev Urol 9: 339–349, 2012. doi:10.1038/nrurol.2012.61.
Crossref | PubMed | Web of Science | Google Scholar - 447. . Biology of human papillomaviruses. Int J Exp Pathol 82: 15–33, 2001. doi:10.1046/j.1365-2613.2001.00177.x.
Crossref | PubMed | Web of Science | Google Scholar - 448. . Detection of human papillomavirus DNA in prostate gland tissue by using the polymerase chain reaction amplification assay. J Clin Microbiol 28: 409–412, 1990. doi:10.1128/JCM.28.3.409-412.1990.
Crossref | PubMed | Web of Science | Google Scholar - 449. . High prevalence of human papillomavirus in prostate tissues. J Urol 145: 850–853, 1991. doi:10.1016/S0022-5347(17)38476-8.
Crossref | PubMed | Web of Science | Google Scholar - 450. . Prevalence of HPV in adults aged 18-69: United States, 2011-2014. NCHS Data Brief 280: 1–8, 2017.
PubMed | Google Scholar - 451. . Foot-and-mouth disease and swine vesicular disease viruses in boar semen. Proc Annu Meet U S Anim Health Assoc 81: 221–230, 1977.
PubMed | Google Scholar - 452. . Zika virus shedding in semen of symptomatic infected men. N Engl J Med 378: 1377–1385, 2018. doi:10.1056/NEJMoa1711038.
Crossref | PubMed | Web of Science | Google Scholar - 453. . Isolation of Aujeszky’s disease virus from boar semen. Acta Vet Acad Sci Hung 29: 29–35, 1981.
PubMed | Google Scholar - 454. . Evidence of chromosomal integration of AAV DNA in human testis tissue. Virus Genes 28: 61–69, 2004. doi:10.1023/B:VIRU.0000012264.54212.f5.
Crossref | PubMed | Web of Science | Google Scholar - 455. . Medical male circumcision and herpes simplex virus 2 acquisition: posttrial surveillance in Kisumu, Kenya. J Infect Dis 208: 1869–1876, 2013. doi:10.1093/infdis/jit371.
Crossref | PubMed | Web of Science | Google Scholar - 456. . Circumcision status and incident herpes simplex virus type 2 infection, genital ulcer disease, and HIV infection. AIDS 26: 1141–1149, 2012. doi:10.1097/QAD.0b013e328352d116.
Crossref | PubMed | Web of Science | Google Scholar - 457. . Microenvironmental signals govern the cellular identity of testicular macrophages. J Leukoc Biol 104: 757–766, 2018. doi:10.1002/JLB.3MR0318-086RR.
Crossref | PubMed | Web of Science | Google Scholar - 458. . Production of the antiviral proteins 2‘5’oligoadenylate synthetase, PKR and Mx in interstitial cells and spermatogonia. J Reprod Immunol 59: 53–60, 2003. doi:10.1016/S0165-0378(02)00061-X.
Crossref | PubMed | Web of Science | Google Scholar - 459. . Sertoli cell-specific deletion of the androgen receptor compromises testicular immune privilege in mice. Biol Reprod 85: 254–260, 2011. doi:10.1095/biolreprod.110.090621.
Crossref | PubMed | Web of Science | Google Scholar - 460. . Androgens regulate the permeability of the blood-testis barrier. Proc Natl Acad Sci USA 102: 16696–16700, 2005. doi:10.1073/pnas.0506084102.
Crossref | PubMed | Web of Science | Google Scholar - 461. . Human viral oncogenesis: a cancer hallmarks analysis. Cell Host Microbe 15: 266–282, 2014. doi:10.1016/j.chom.2014.02.011.
Crossref | PubMed | Web of Science | Google Scholar - 462. . Endogenous retroviruses: with us and against us. Front Chem 5: 23, 2017. doi:10.3389/fchem.2017.00023.
Crossref | PubMed | Web of Science | Google Scholar - 463. . Epididymitis: revelations at the convergence of clinical and basic sciences. Asian J Androl 17: 756–763, 2015.
PubMed | Web of Science | Google Scholar - 464. . Diffuse lymphoid tissue associated with the human bulbourethral gland. An immunohistologic characterization. J Androl 13: 337–341, 1992.
PubMed | Google Scholar - 465. . Pathology and localization of simian immunodeficiency virus in the reproductive tract of chronically infected male rhesus macaques. Lab Invest 70: 255–262, 1994.
PubMed | Web of Science | Google Scholar - 466. . In situ detection of regulatory T cells in human genital herpes simplex virus type 2 (HSV-2) reactivation and their influence on spontaneous HSV-2 reactivation. J Infect Dis 214: 23–31, 2016. doi:10.1093/infdis/jiw091.
Crossref | PubMed | Web of Science | Google Scholar - 467. . Effect of intratesticular inoculation with Aujeszky’s disease virus on genital organs of boars. Vet Microbiol 14: 355–363, 1987. doi:10.1016/0378-1135(87)90027-7.
Crossref | PubMed | Web of Science | Google Scholar - 468. . Transposable elements and their epigenetic regulation in mental disorders: current evidence in the field. Front Genet 10: 580, 2019. doi:10.3389/fgene.2019.00580.
Crossref | PubMed | Web of Science | Google Scholar - 469. . The blood-testis and blood-epididymis barriers are more than just their tight junctions. Biol Reprod 84: 851–858, 2011. doi:10.1095/biolreprod.110.087452.
Crossref | PubMed | Web of Science | Google Scholar - 470. . Maternal alloantigens promote the development of tolerogenic fetal regulatory T cells in utero. Science 322: 1562–1565, 2008. doi:10.1126/science.1164511.
Crossref | PubMed | Web of Science | Google Scholar - 471. . Asymptomatic seminal infection of herpes simplex virus: impact on male infertility. J Biomed Res 27: 56–61, 2013. doi:10.7555/JBR.27.20110139.
Crossref | PubMed | Google Scholar - 472. . Latent BK virus infection and Kaposi’s sarcoma pathogenesis. Int J Cancer 66: 717–722, 1996. doi:10.1002/(SICI)1097-0215(19960611)66:6<717:AID-IJC1>3.0.CO;2-2.
Crossref | PubMed | Web of Science | Google Scholar - 473. . Detection of human herpesvirus 8 (HHV-8) in normal prostates. Prostate 66: 1302–1310, 2006. doi:10.1002/pros.20459.
Crossref | PubMed | Web of Science | Google Scholar - 474. . Impact of short-term HAART initiated during the chronic stage or shortly post-exposure on SIV infection of male genital organs. PLoS One 7:
e37348 , 2012. doi:10.1371/journal.pone.0037348.
Crossref | PubMed | Web of Science | Google Scholar - 475. . Sexually acquired Zika virus: a systematic review. Clin Microbiol Infect 23: 296–305, 2017. doi:10.1016/j.cmi.2016.12.027.
Crossref | PubMed | Web of Science | Google Scholar - 476. . History of mosquitoborne diseases in the United States and implications for new pathogens. Emerg Infect Dis 24: 821–826, 2018. doi:10.3201/eid2405.171609.
Crossref | PubMed | Web of Science | Google Scholar - 477. . Sperm ultrastructure and meiotic segregation in a group of patients with chronic hepatitis B and C. Andrologia 40: 286–291, 2008. doi:10.1111/j.1439-0272.2008.00855.x.
Crossref | PubMed | Google Scholar - 478. . Infectious burden and semen parameters. Urology 100: 90–96, 2017. doi:10.1016/j.urology.2016.10.032.
Crossref | PubMed | Web of Science | Google Scholar - 479. . Transmission of chromosomally integrated human herpesvirsus 6 (HHV-6) variant A from a parent to children leading to misdiagnosis of active HHV-6 infection. Transpl Infect Dis 11: 503–506, 2009. doi:10.1111/j.1399-3062.2009.00430.x.
Crossref | PubMed | Web of Science | Google Scholar - 480. . Herpesviruses and chromosomal integration. J Virol 84: 12100–12109, 2010. doi:10.1128/JVI.01169-10.
Crossref | PubMed | Web of Science | Google Scholar - 481. . Estimation of country-specific and global prevalence of male circumcision. [Correction in Popul Health Metr 14: 11, 2016.] Popul Health Metr 14: 4, 2016. doi:10.1186/s12963-016-0073-5.
Crossref | PubMed | Web of Science | Google Scholar - 482. . Longitudinal viral dynamics in semen during early HIV infection. Clin Infect Dis 64: 428–434, 2017. doi:10.1093/cid/ciw784.
Crossref | PubMed | Web of Science | Google Scholar - 483. . Animal models of Zika virus infection, pathogenesis, and immunity. J Virol 91:
e00009-17 , 2017. doi:10.1128/JVI.00009-17.
Crossref | PubMed | Web of Science | Google Scholar - 484. . Developmental origin and maintenance of distinct testicular macrophage populations. J Exp Med 214: 2829–2841, 2017. doi:10.1084/jem.20170829.
Crossref | PubMed | Web of Science | Google Scholar - 485. . Testicular macrophages: guardians of fertility. Cell Immunol 330: 120–125, 2018. doi:10.1016/j.cellimm.2018.03.009.
Crossref | PubMed | Web of Science | Google Scholar - 486. . Testicular germ cells of HIV-seropositive asymptomatic men are infected by the virus. J Reprod Immunol 41: 81–93, 1998. doi:10.1016/S0165-0378(98)00050-3.
Crossref | PubMed | Web of Science | Google Scholar - 487. . Semen inhibits Zika virus infection of cells and tissues from the anogenital region. Nat Commun 9: 2207, 2018. doi:10.1038/s41467-018-04442-y.
Crossref | PubMed | Web of Science | Google Scholar - 488. . Semen-derived amyloid fibrils drastically enhance HIV infection. Cell 131: 1059–1071, 2007. doi:10.1016/j.cell.2007.10.014.
Crossref | PubMed | Web of Science | Google Scholar - 489. . Effect of semen and seminal amyloid on vaginal transmission of simian immunodeficiency virus. Retrovirology 10: 148, 2013. doi:10.1186/1742-4690-10-148.
Crossref | PubMed | Web of Science | Google Scholar - 490. . Emerging viral infections and their impact on the global burden of neurological disease. Semin Neurol 38: 163–175, 2018. doi:10.1055/s-0038-1647247.
Crossref | PubMed | Web of Science | Google Scholar - 491. . Sexual transmission of human T-lymphotropic virus type I (HTLV-I). Ann Intern Med 111: 555–560, 1989. doi:10.7326/0003-4819-111-7-555.
Crossref | PubMed | Web of Science | Google Scholar - 492. . Yellow fever: the Pacific should be prepared. Lancet 392: 2347, 2018. doi:10.1016/S0140-6736(18)32520-0.
Crossref | PubMed | Web of Science | Google Scholar - 493. . Potential sexual transmission of Zika virus. Emerg Infect Dis 21: 359–361, 2015. doi:10.3201/eid2102.141363.
Crossref | PubMed | Web of Science | Google Scholar - 494. . Oncogenic herpesvirus HHV-8 promotes androgen-independent prostate cancer growth. Cancer Res 73: 5695–5708, 2013. doi:10.1158/0008-5472.CAN-12-4196.
Crossref | PubMed | Web of Science | Google Scholar - 495. . Bovine immunodeficiency virus in stud bull semen. Am J Vet Res 56: 760–763, 1995.
PubMed | Web of Science | Google Scholar - 496. . Presence and distribution of leucocyte subsets in the murine epididymis after vasectomy. Int J Androl 13: 39–49, 1990. doi:10.1111/j.1365-2605.1990.tb00958.x.
Crossref | PubMed | Google Scholar - 497. . Immuno-competent cells in the murine epididymis. Int J Androl 12: 85–94, 1989. doi:10.1111/j.1365-2605.1989.tb01289.x.
Crossref | PubMed | Google Scholar - 498. . Epidemiology. In: Fields Virology, edited by Knipe D, Howley P. Philadelphia, PA: Lippincott Williams & Wilkins, 2013.
Google Scholar - 499. . Detection and quantification of human herpes viruses types 4-6 in sperm samples of patients with fertility disorders and chronic inflammatory urogenital tract diseases. Andrology 2: 687–694, 2014. doi:10.1111/j.2047-2927.2014.00232.x.
Crossref | PubMed | Web of Science | Google Scholar - 500. . Detection of human cytomegalovirus in motile spermatozoa and spermatogenic cells in testis organotypic culture. Herpesviridae 2: 7, 2011. doi:10.1186/2042-4280-2-7.
Crossref | PubMed | Google Scholar - 501. . Exposure to HIV-1 directly impairs mucosal epithelial barrier integrity allowing microbial translocation. PLoS Pathog 6:
e1000852 , 2010. doi:10.1371/journal.ppat.1000852.
Crossref | PubMed | Web of Science | Google Scholar - 502. . Epidemiology of hepatitis B virus infection and impact of vaccination on disease. Clin Liver Dis 20: 607–628, 2016. doi:10.1016/j.cld.2016.06.006.
Crossref | PubMed | Web of Science | Google Scholar - 503. . Prevalence of human herpes virus types 1-7 in the semen of men attending an infertility clinic and correlation with semen parameters. Fertil Steril 91: 2487–2494, 2009. doi:10.1016/j.fertnstert.2008.03.074.
Crossref | PubMed | Web of Science | Google Scholar - 504. . Laboratory diagnosis and transmissibility of bovine viral diarrhea virus from a bull with a persistent testicular infection. Vet Microbiol 170: 246–257, 2014. doi:10.1016/j.vetmic.2014.02.028.
Crossref | PubMed | Web of Science | Google Scholar - 505. . Innate and adaptive immune responses in male and female reproductive tracts in homeostasis and following HIV infection. Cell Mol Immunol 11: 410–427, 2014. doi:10.1038/cmi.2014.41.
Crossref | PubMed | Web of Science | Google Scholar - 506. . Persistent detection of Zika virus RNA in semen for six months after symptom onset in a traveller returning from Haiti to Italy, February 2016. Euro Surveill 21: 30314, 2016. doi:10.2807/1560-7917.ES.2016.21.32.30314.
Crossref | PubMed | Google Scholar - 507. . Consensus development of a histopathological classification system for chronic prostatic inflammation. BJU Int 87: 797–805, 2001. doi:10.1046/j.1464-410x.2001.02193.x.
Crossref | PubMed | Web of Science | Google Scholar - 508. Nieschlag E, Behre HM (Editors). Andrology (3rd ed.). Berlin: Springer, 2010.
Crossref | Google Scholar - 509. . [A case of a laboratory infection with Marburg fever]. Zh Mikrobiol Epidemiol Immunobiol 3: 104–106, 1994.
PubMed | Google Scholar - 510. . Tissue-specific mRNA expression profiles of human toll-like receptors and related genes. Biol Pharm Bull 28: 886–892, 2005. doi:10.1248/bpb.28.886.
Crossref | PubMed | Web of Science | Google Scholar - 511. . HIV-1 nucleic acids localize to the spermatogonia and their progeny. A study by polymerase chain reaction in situ hybridization. Am J Pathol 144: 1142–1148, 1994.
PubMed | Web of Science | Google Scholar - 512. . A role for surfactant protein D in innate immunity of the human prostate. Prostate 65: 241–251, 2005. doi:10.1002/pros.20292.
Crossref | PubMed | Web of Science | Google Scholar - 513. . Regulation of surfactant protein D in the rodent prostate. Reprod Biol Endocrinol 5: 42, 2007. doi:10.1186/1477-7827-5-42.
Crossref | PubMed | Web of Science | Google Scholar - 514. . Lesions and cellular tropism of natural rift valley fever virus infection in adult sheep. Vet Pathol 56: 61–77, 2019. doi:10.1177/0300985818806049.
Crossref | PubMed | Web of Science | Google Scholar - 515. . In utero priming of highly functional effector T cell responses to human malaria. Sci Transl Med 10:
eaat6176 , 2018.
Crossref | PubMed | Web of Science | Google Scholar - 516. . Adverse effects of hepatitis B virus on sperm motility and fertilization ability during IVF. Reprod Biomed Online 23: 207–212, 2011. doi:10.1016/j.rbmo.2011.04.008.
Crossref | PubMed | Web of Science | Google Scholar - 517. . Secretory leucocyte protease inhibitor in the male genital tract: PSA-induced proteolytic processing in human semen and tissue localization. J Androl 16: 64–74, 1995.
PubMed | Google Scholar - 518. . Bovine herpesvirus 1 in semen of bulls and the risk of transmission: a brief review. Vet Q 17: 29–33, 1995. doi:10.1080/01652176.1995.9694526.
Crossref | PubMed | Web of Science | Google Scholar - 519. . Persistence and intra-host genetic evolution of zika virus infection in symptomatic adults: a special view in the male reproductive system. Viruses 10: 615, 2018. doi:10.3390/v10110615.
Crossref | PubMed | Web of Science | Google Scholar - 520. . Urinary and sexual manifestations of patients infected by HTLV-I. Clinics (São Paulo) 62: 191–196, 2007. doi:10.1590/S1807-59322007000200015.
Crossref | PubMed | Web of Science | Google Scholar - 521. . Prevalence of erectile dysfunction in HTLV-1-infected patients and its association with overactive bladder. Urology 75: 1100–1103, 2010. doi:10.1016/j.urology.2009.11.041.
Crossref | PubMed | Web of Science | Google Scholar - 522. . Distinct cytokine patterns in semen influence local HIV shedding and HIV target cell activation. J Infect Dis 209: 1174–1184, 2014. doi:10.1093/infdis/jit649.
Crossref | PubMed | Web of Science | Google Scholar - 523. . Clinical and mucosal immune correlates of HIV-1 semen levels in antiretroviral-naive men. Open Forum Infect Dis 4:
ofx033 , 2017. doi:10.1093/ofid/ofx033.
Crossref | PubMed | Web of Science | Google Scholar - 524. . Zika viral dynamics and shedding in rhesus and cynomolgus macaques. Nat Med 22: 1448–1455, 2016. doi:10.1038/nm.4206.
Crossref | PubMed | Web of Science | Google Scholar - 525. . HERVd: the human endogenous retroviruses database: update. Nucleic Acids Res 32: 50D, 2004. doi:10.1093/nar/gkh075.
Crossref | PubMed | Web of Science | Google Scholar - 526. . Sexual transmission of human T-cell lymphotropic virus type 1. Rev Soc Bras Med Trop 47: 265–274, 2014. doi:10.1590/0037-8682-0232-2013.
Crossref | PubMed | Web of Science | Google Scholar - 527. . Members of the Toll-like receptor family of innate immunity pattern-recognition receptors are abundant in the male rat reproductive tract. Biol Reprod 76: 958–964, 2007. doi:10.1095/biolreprod.106.059410.
Crossref | PubMed | Web of Science | Google Scholar - 528. . Localization of Toll-like receptors on epididymal epithelial cells and spermatozoa. Am J Reprod Immunol 60: 541–555, 2008. doi:10.1111/j.1600-0897.2008.00654.x.
Crossref | PubMed | Web of Science | Google Scholar - 529. . Herpesvirus, cytomegalovirus, human sperm and assisted fertilization. Hum Reprod 17: 1281–1287, 2002. doi:10.1093/humrep/17.5.1281.
Crossref | PubMed | Web of Science | Google Scholar - 530. . Low frequency of male circumcision and unwillingness to be circumcised among MSM in Buenos Aires, Argentina: association with sexually transmitted infections. J Int AIDS Soc 16: 18500, 2013. doi:10.7448/IAS.16.1.18500.
Crossref | PubMed | Web of Science | Google Scholar - 531. . Recombinant origin of the retrovirus XMRV. Science 333: 97–101, 2011. doi:10.1126/science.1205292.
Crossref | PubMed | Web of Science | Google Scholar - 532. . Intermittent detection of hepatitis C virus (HCV) in semen from men with human immunodeficiency virus type 1 (HIV-1) and HCV. J Med Virol 69: 344–349, 2003. doi:10.1002/jmv.10295.
Crossref | PubMed | Web of Science | Google Scholar - 533. . Patterns of residual HIV-1 RNA shedding in the seminal plasma of patients on effective antiretroviral therapy. Basic Clin Androl 27: 17, 2017. doi:10.1186/s12610-017-0063-x.
Crossref | PubMed | Web of Science | Google Scholar - 534. . Estimating per-act HIV transmission risk: a systematic review. AIDS 28: 1509–1519, 2014. doi:10.1097/QAD.0000000000000298.
Crossref | PubMed | Web of Science | Google Scholar - 535. . Susceptibility to human immunodeficiency virus-1 infection of human foreskin and cervical tissue grown in explant culture. Am J Pathol 161: 867–873, 2002. doi:10.1016/S0002-9440(10)64247-2.
Crossref | PubMed | Web of Science | Google Scholar - 536. . Persistence of Zika virus in body fluids - final report. N Engl J Med 380: 198–199, 2019. doi:10.1056/NEJMc1814416.
Crossref | PubMed | Web of Science | Google Scholar - 537. . Human benign prostatic hyperplasia stromal cells as inducers and targets of chronic immuno-mediated inflammation. J Immunol 182: 4056–4064, 2009. doi:10.4049/jimmunol.0801875.
Crossref | PubMed | Web of Science | Google Scholar - 538. . Dual role of immune cells in the testis: protective or pathogenic for germ cells? Spermatogenesis 3:
e23870 , 2013. doi:10.4161/spmg.23870.
Crossref | PubMed | Google Scholar - 539. . Human papillomavirus infection in couples undergoing in vitro fertilization procedures: impact on reproductive outcomes. Fertil Steril 95: 1845–1848, 2011. doi:10.1016/j.fertnstert.2010.11.047.
Crossref | PubMed | Web of Science | Google Scholar - 540. . The epidemiology of prostate cancer. Cold Spring Harb Perspect Med 8:
a030361 , 2018. doi:10.1101/cshperspect.a030361.
Crossref | PubMed | Web of Science | Google Scholar - 541. . Ebola virus localization in the Macaque reproductive tract during acute Ebola virus disease. Am J Pathol 188: 550–558, 2018. doi:10.1016/j.ajpath.2017.11.004.
Crossref | PubMed | Web of Science | Google Scholar - 542. . BVDV: a pestivirus inducing tolerance of the innate immune response. Biologicals 41: 39–51, 2013. doi:10.1016/j.biologicals.2012.07.006.
Crossref | PubMed | Web of Science | Google Scholar - 543. . Presence of pro-lentiviral DNA in male sexual organs and ejaculates of small ruminants. Theriogenology 69: 433–442, 2008. doi:10.1016/j.theriogenology.2007.10.013.
Crossref | PubMed | Web of Science | Google Scholar - 544. . Isolation and characterization of viruses from semen and the reproductive tract of male swine. J Am Vet Med Assoc 161: 1306–1316, 1972.
PubMed | Web of Science | Google Scholar - 545. . Semen quality in HIV patients under stable antiretroviral therapy is impaired compared to WHO 2010 reference values and on sperm proteome level. AIDS 28: 875–880, 2014. doi:10.1097/QAD.0000000000000161.
Crossref | PubMed | Web of Science | Google Scholar - 546. . Transmission between archaic and modern human ancestors during the evolution of the oncogenic human papillomavirus 16. Mol Biol Evol 34: 4–19, 2017. doi:10.1093/molbev/msw214.
Crossref | PubMed | Web of Science | Google Scholar - 547. . HIV-1 is undetectable in preejaculatory secretions from HIV-1-infected men on suppressive HAART. AIDS 30: 1899–1903, 2016. doi:10.1097/QAD.0000000000001130.
Crossref | PubMed | Web of Science | Google Scholar - 548. . Highly active antiretroviral therapy does not completely suppress HIV in semen of sexually active HIV-infected men who have sex with men. AIDS 26: 1535–1543, 2012. doi:10.1097/QAD.0b013e328353b11b.
Crossref | PubMed | Web of Science | Google Scholar - 549. . Evidence of excretion of Schmallenberg virus in bull semen. Vet Res (Faisalabad) 45: 37, 2014. doi:10.1186/1297-9716-45-37.
Crossref | PubMed | Web of Science | Google Scholar - 550. . Characterization of myeloid cell populations in human testes collected after sex reassignment surgery. J Reprod Immunol 125: 16–24, 2018. doi:10.1016/j.jri.2017.10.043.
Crossref | PubMed | Web of Science | Google Scholar - 551. . Distinct defensin profiles in Neisseria gonorrhoeae and Chlamydia trachomatis urethritis reveal novel epithelial cell-neutrophil interactions. Infect Immun 73: 4823–4833, 2005. doi:10.1128/IAI.73.8.4823-4833.2005.
Crossref | PubMed | Web of Science | Google Scholar - 552. . Clinical and epidemiological aspects of HTLV-II infection in São Paulo, Brazil: presence of tropical spastic paraparesis/HTLV-associated myelopathy (TSP/HAM) simile diagnosis in HIV-1-co-infected subjects. Rev Inst Med Trop São Paulo 48: 207–210, 2006. doi:10.1590/S0036-46652006000400006.
Crossref | PubMed | Google Scholar - 553. . The effects of circumcision on the penis microbiome. PLoS One 5:
e8422 , 2010. doi:10.1371/journal.pone.0008422.
Crossref | PubMed | Web of Science | Google Scholar - 554. . Porcine reproductive and respiratory syndrome virus infection in the boar: a review. Theriogenology 63: 1–16, 2005. doi:10.1016/j.theriogenology.2004.03.018.
Crossref | PubMed | Web of Science | Google Scholar - 555. . Foreskin T-cell subsets differ substantially from blood with respect to HIV co-receptor expression, inflammatory profile, and memory status. Mucosal Immunol 5: 121–128, 2012. doi:10.1038/mi.2011.56.
Crossref | PubMed | Web of Science | Google Scholar - 556. . Chemokine levels in the penile coronal sulcus correlate with HIV-1 acquisition and are reduced by male circumcision in Rakai, Uganda. PLoS Pathog 12:
e1006025 , 2016. doi:10.1371/journal.ppat.1006025.
Crossref | PubMed | Web of Science | Google Scholar - 557. . Orchitis and human immunodeficiency virus type 1 infected cells in reproductive tissues from men with the acquired immune deficiency syndrome. Am J Pathol 139: 149–160, 1991.
PubMed | Web of Science | Google Scholar - 558. . Innate and acquired immunity in the human penile urethra. J Reprod Immunol 88: 219–227, 2011. doi:10.1016/j.jri.2011.01.006.
Crossref | PubMed | Web of Science | Google Scholar - 559. . Expression of toll-like receptors in genital tract tissues from normal and HIV-infected men. Am J Reprod Immunol 65: 28–43, 2011. doi:10.1111/j.1600-0897.2010.00877.x.
Crossref | PubMed | Web of Science | Google Scholar - 560. . Testicular degeneration and infertility following arbovirus infection. J Virol 92:
e01131-18 , 2018. doi:10.1128/JVI.01131-18.
Crossref | PubMed | Web of Science | Google Scholar - 561. . International patterns and trends in testis cancer incidence. Int J Cancer 115: 822–827, 2005. doi:10.1002/ijc.20931.
Crossref | PubMed | Web of Science | Google Scholar - 562. . Effect of hepatitis B virus infection on sperm quality and oxidative stress state of the semen of infertile males. Am J Reprod Immunol 76: 183–185, 2016. doi:10.1111/aji.12537.
Crossref | PubMed | Web of Science | Google Scholar - 563. . Androgen depletion augments antibacterial prostate host defences in rats. Int J Androl 35: 845–859, 2012. doi:10.1111/j.1365-2605.2012.01288.x.
Crossref | PubMed | Google Scholar - 564. . Androgen regulation of host defenses and response to inflammatory stimuli in the prostate gland. Cell Biol Int 41: 1223–1233, 2017. doi:10.1002/cbin.10755.
Crossref | PubMed | Web of Science | Google Scholar - 565. . Toll-like receptor 4 in rat prostate: modulation by testosterone and acute bacterial infection in epithelial and stromal cells. Biol Reprod 75: 664–672, 2006. doi:10.1095/biolreprod.106.053967.
Crossref | PubMed | Web of Science | Google Scholar - 566. . Favipiravir and ribavirin treatment of epidemiologically linked cases of Lassa fever. Clin Infect Dis 65: 855–859, 2017. doi:10.1093/cid/cix406.
Crossref | PubMed | Web of Science | Google Scholar - 567. . The evolving field of human papillomavirus receptor research: a review of binding and entry. J Virol 87: 6062–6072, 2013. doi:10.1128/JVI.00330-13.
Crossref | PubMed | Web of Science | Google Scholar - 568. . Penile appearance, lumps and bumps. Aust Fam Physician 42: 270–274, 2013.
PubMed | Google Scholar - 569. . Metagenomics and the human virome in asymptomatic individuals. Annu Rev Microbiol 70: 125–141, 2016. doi:10.1146/annurev-micro-102215-095431.
Crossref | PubMed | Web of Science | Google Scholar - 570. . Genetic evidence that captured retroviral envelope syncytins contribute to myoblast fusion and muscle sexual dimorphism in mice. PLoS Genet 12:
e1006289 , 2016. doi:10.1371/journal.pgen.1006289.
Crossref | PubMed | Web of Science | Google Scholar - 572. . Sertoli cells initiate testicular innate immune responses through TLR activation. J Immunol 177: 7122–7130, 2006. doi:10.4049/jimmunol.177.10.7122.
Crossref | PubMed | Web of Science | Google Scholar - 573. . Human papillomavirus DNA is found in the vas deferens. J Infect Dis 185: 1664–1667, 2002. doi:10.1086/340421.
Crossref | PubMed | Web of Science | Google Scholar - 574. . Natural co-infection of torque teno virus and porcine circovirus 2 in the reproductive apparatus of swine. Res Vet Sci 92: 519–523, 2012. doi:10.1016/j.rvsc.2011.04.001.
Crossref | PubMed | Web of Science | Google Scholar - 575. . Persistence of porcine rubulavirus in experimentally infected boars. Vet Microbiol 162: 491–498, 2013. doi:10.1016/j.vetmic.2012.10.037.
Crossref | PubMed | Web of Science | Google Scholar - 576. . Interaction of fibronectin with semen amyloids synergistically enhances HIV infection. J Infect Dis 210: 1062–1066, 2014. doi:10.1093/infdis/jiu220.
Crossref | PubMed | Web of Science | Google Scholar - 577. . Liquefaction of semen generates and later degrades a conserved semenogelin peptide that enhances HIV infection. J Virol 88: 7221–7234, 2014. doi:10.1128/JVI.00269-14.
Crossref | PubMed | Web of Science | Google Scholar - 578. . Peptides released by physiological cleavage of semen coagulum proteins form amyloids that enhance HIV infection. Cell Host Microbe 10: 541–550, 2011. doi:10.1016/j.chom.2011.10.010.
Crossref | PubMed | Web of Science | Google Scholar - 579. . Genital herpes in young adults: changing sexual behaviours, epidemiology and management. Herpes 12: 10–14, 2005.
PubMed | Google Scholar - 580. . Seminal fluid and fertility in women. Fertil Steril 106: 511–519, 2016. doi:10.1016/j.fertnstert.2016.07.1101.
Crossref | PubMed | Web of Science | Google Scholar - 581. . Male germ cells support long-term propagation of Zika virus. Nat Commun 9: 2090, 2018. doi:10.1038/s41467-018-04444-w.
Crossref | PubMed | Web of Science | Google Scholar - 582. . Myeloid cell interaction with HIV: a complex relationship. Front Immunol 8: 1698, 2017. doi:10.3389/fimmu.2017.01698.
Crossref | PubMed | Web of Science | Google Scholar - 583. . Persistence and genetic stability of Ebola virus during the outbreak in Kikwit, Democratic Republic of the Congo, 1995. J Infect Dis 179, Suppl 1: S170–S176, 1999. doi:10.1086/514291.
Crossref | PubMed | Web of Science | Google Scholar - 584. . Herpesviridae. In: Fields Virology, edited by Knipe DM, Howley PM. Philadelphia, PA: Lippincott Williams & Wilkins, 2013.
Google Scholar - 585. . Retroviral infection in Peruvian men who have sex with men. Clin Infect Dis 49: 112–117, 2009. doi:10.1086/599609.
Crossref | PubMed | Web of Science | Google Scholar - 586. . Characterisation of macaque testicular leucocyte populations and T-lymphocyte immunity. J Reprod Immunol 100: 146–156, 2013. doi:10.1016/j.jri.2013.09.003.
Crossref | PubMed | Web of Science | Google Scholar - 587. . Prevalence and incidence of Zika virus infection among household contacts of patients with Zika virus disease, Puerto Rico, 2016–2017. J Infect Dis 220: 932–939, 2019.
PubMed | Web of Science | Google Scholar - 588. . Antiviral antibodies and T cells are present in the foreskin of simian immunodeficiency virus-infected rhesus macaques. J Virol 86: 7098–7106, 2012. doi:10.1128/JVI.00410-12.
Crossref | PubMed | Web of Science | Google Scholar - 589. . The epidemiology and disease outcomes of human T-lymphotropic virus type II. AIDS Rev 6: 144–154, 2004.
PubMed | Web of Science | Google Scholar - 590. . Susceptibility of human testis to human immunodeficiency virus-1 infection in situ and in vitro. Am J Pathol 169: 2094–2103, 2006. doi:10.2353/ajpath.2006.060191.
Crossref | PubMed | Web of Science | Google Scholar - 591. . Clinical, virologic, and immunologic follow-up of convalescent Ebola hemorrhagic fever patients and their household contacts, Kikwit, Democratic Republic of the Congo. Commission de Lutte contre les Epidémies à Kikwit. J Infect Dis 179, Suppl 1: S28–S35, 1999. doi:10.1086/514318.
Crossref | PubMed | Web of Science | Google Scholar - 592. . Dynamic control of endogenous retroviruses during development. Virology 411: 273–287, 2011. doi:10.1016/j.virol.2010.12.007.
Crossref | PubMed | Web of Science | Google Scholar - 593. . Sexual transmission of HIV. N Engl J Med 336: 1072–1078, 1997. doi:10.1056/NEJM199704103361507.
Crossref | PubMed | Web of Science | Google Scholar - 594. . Molecular biology, pathogenesis and pathology of mumps virus. J Pathol 235: 242–252, 2015. doi:10.1002/path.4445.
Crossref | PubMed | Web of Science | Google Scholar - 595. . Infectious pathogens potentially transmitted by semen of the black variety of the Manchega sheep breed: Health constraints for conservation purposes. Anim Reprod Sci 149: 152–157, 2014. doi:10.1016/j.anireprosci.2014.07.006.
Crossref | PubMed | Web of Science | Google Scholar - 596. . Mucin gene expression in human male urogenital tract epithelia. Hum Reprod 21: 2783–2793, 2006. doi:10.1093/humrep/del164.
Crossref | PubMed | Web of Science | Google Scholar - 597. . Semen clusterin is a novel DC-SIGN ligand. J Immunol 187: 5299–5309, 2011. doi:10.4049/jimmunol.1101889.
Crossref | PubMed | Web of Science | Google Scholar - 598. . Community-based prevalence of genital human papilloma virus (HPV) infection: a systematic review and meta-analysis. Asian Pac J Cancer Prev 18: 145–154, 2017. doi:10.22034/APJCP.2017.18.1.145.
Crossref | PubMed | Google Scholar - 599. . Transforming growth factor beta 1 up-regulates CD169 (sialoadhesin) expression on monocyte-derived dendritic cells: role in HIV sexual transmission. AIDS 28: 2375–2380, 2014. doi:10.1097/QAD.0000000000000431.
Crossref | PubMed | Web of Science | Google Scholar - 600. . High prevalence of human cytomegalovirus in prostatic intraepithelial neoplasia and prostatic carcinoma. J Urol 170: 998–1002, 2003. doi:10.1097/01.ju.0000080263.46164.97.
Crossref | PubMed | Web of Science | Google Scholar - 601. . Targeted integration of adeno-associated virus (AAV) into human chromosome 19. EMBO J 10: 3941–3950, 1991. doi:10.1002/j.1460-2075.1991.tb04964.x.
Crossref | PubMed | Web of Science | Google Scholar - 602. . Excess type I interferon signaling in the mouse seminiferous tubules leads to germ cell loss and sterility. J Biol Chem 286: 23280–23295, 2011. doi:10.1074/jbc.M111.229120.
Crossref | PubMed | Web of Science | Google Scholar - 603. . Avoiding the void: cell-to-cell spread of human viruses. Nat Rev Microbiol 6: 815–826, 2008. doi:10.1038/nrmicro1972.
Crossref | PubMed | Web of Science | Google Scholar - 604. . Herpes genitalis: diagnosis, treatment and prevention. Geburtshilfe Frauenheilkd 76: 1310–1317, 2016. doi:10.1055/s-0042-116494.
Crossref | PubMed | Web of Science | Google Scholar - 605. . Human microbiome and HIV/AIDS. Curr HIV/AIDS Rep 9: 44–51, 2012. doi:10.1007/s11904-011-0103-7.
Crossref | PubMed | Web of Science | Google Scholar - 606. . Frequent recovery of HIV-1 from genital herpes simplex virus lesions in HIV-1-infected men. JAMA 280: 61–66, 1998. doi:10.1001/jama.280.1.61.
Crossref | PubMed | Web of Science | Google Scholar - 607. . Herpes simplex virus type 1 exhibits a tropism for basal entry in polarized epithelial cells. J Gen Virol 84: 2473–2484, 2003. doi:10.1099/vir.0.19226-0.
Crossref | PubMed | Web of Science | Google Scholar - 608. . Rapid host immune response and viral dynamics in herpes simplex virus-2 infection. Nat Med 19: 280–288, 2013. doi:10.1038/nm.3103.
Crossref | PubMed | Web of Science | Google Scholar - 609. . Persistence and sexual transmission of filoviruses. Viruses 10: 683, 2018. doi:10.3390/v10120683.
Crossref | PubMed | Web of Science | Google Scholar - 610. . Transcriptional profiling of human endogenous retrovirus group HERV-K(HML-2) loci in melanoma. Genome Biol Evol 5: 307–328, 2013. doi:10.1093/gbe/evt010.
Crossref | PubMed | Web of Science | Google Scholar - 611. . Interferon-stimulated genes and their antiviral effector functions. Curr Opin Virol 1: 519–525, 2011. doi:10.1016/j.coviro.2011.10.008.
Crossref | PubMed | Web of Science | Google Scholar - 612. . Risk factors for human T-cell lymphotropic virus types I and II (HTLV-I and -II) in blood donors: the Retrovirus Epidemiology Donor Study. NHLBI Retrovirus Epidemiology Donor Study. J Acquir Immune Defic Syndr Hum Retrovirol 14: 263–271, 1997. doi:10.1097/00042560-199703010-00011.
Crossref | PubMed | Google Scholar - 613. . Chronic orchitis: a neglected cause of male infertility? Andrologia 40: 84–91, 2008. doi:10.1111/j.1439-0272.2008.00837.x.
Crossref | PubMed | Web of Science | Google Scholar - 614. . Congenital infection with atypical porcine pestivirus (APPV) is associated with disease and viral persistence. Vet Res (Faisalabad) 48: 1, 2017. doi:10.1186/s13567-016-0406-1.
Crossref | PubMed | Web of Science | Google Scholar - 615. . A reverse transcriptase-dependent mechanism is essential for murine preimplantation development. Genes (Basel) 2: 360–373, 2011. doi:10.3390/genes2020360.
Crossref | PubMed | Web of Science | Google Scholar - 616. . The impact of drugs on male fertility: a review. Andrology 5: 640–663, 2017. doi:10.1111/andr.12366.
Crossref | PubMed | Web of Science | Google Scholar - 617. . Absence of hepatitis C virus and detection of hepatitis G virus/GB virus C RNA sequences in the semen of infected men. J Infect Dis 177: 848–854, 1998. doi:10.1086/515257.
Crossref | PubMed | Web of Science | Google Scholar - 618. . The human penis is a genuine immunological effector site. Front Immunol 8: 1732, 2017. doi:10.3389/fimmu.2017.01732.
Crossref | PubMed | Web of Science | Google Scholar - 619. . High risk of temporary alteration of semen parameters after recent acute febrile illness. Fertil Steril 88: 970.e1–970.e7, 2007. doi:10.1016/j.fertnstert.2006.12.045.
Crossref | PubMed | Web of Science | Google Scholar - 620. . Toll-like receptor-initiated testicular innate immune responses in mouse Leydig cells. Endocrinology 152: 2827–2836, 2011. doi:10.1210/en.2011-0031.
Crossref | PubMed | Web of Science | Google Scholar - 621. . Autophagy regulates spermatid differentiation via degradation of PDLIM1. Autophagy 12: 1575–1592, 2016. doi:10.1080/15548627.2016.1192750.
Crossref | PubMed | Web of Science | Google Scholar - 622. . TGF-β mediates proinflammatory seminal fluid signaling in human cervical epithelial cells. J Immunol 189: 1024–1035, 2012. doi:10.4049/jimmunol.1200005.
Crossref | PubMed | Web of Science | Google Scholar - 623. . Seminal fluid induces leukocyte recruitment and cytokine and chemokine mRNA expression in the human cervix after coitus. J Immunol 188: 2445–2454, 2012. doi:10.4049/jimmunol.1102736.
Crossref | PubMed | Web of Science | Google Scholar - 624. . Best practice policies for male infertility. Fertil Steril 77: 873–882, 2002. doi:10.1016/S0015-0282(02)03105-9.
Crossref | PubMed | Web of Science | Google Scholar - 625. . The testis and epididymis are productively infected by SIV and SHIV in juvenile macaques during the post-acute stage of infection. Retrovirology 4: 7, 2007. doi:10.1186/1742-4690-4-7.
Crossref | PubMed | Web of Science | Google Scholar - 626. . Interactions between HIV-1 and mucosal cells in the female reproductive tract. Am J Reprod Immunol 71: 608–617, 2014. doi:10.1111/aji.12244.
Crossref | PubMed | Web of Science | Google Scholar - 627. . Sertoli cells are susceptible to ZIKV infection in mouse testis. Front Cell Infect Microbiol 7: 272, 2017. doi:10.3389/fcimb.2017.00272.
Crossref | PubMed | Web of Science | Google Scholar - 628. . Disproportionately high semen shedding of HIV is associated with compartmentalized cytomegalovirus reactivation. J Infect Dis 193: 45–48, 2006. doi:10.1086/498576.
Crossref | PubMed | Web of Science | Google Scholar - 629. . Persistent HIV RNA shedding in semen despite effective antiretroviral therapy. AIDS 23: 2050–2054, 2009. doi:10.1097/QAD.0b013e3283303e04.
Crossref | PubMed | Web of Science | Google Scholar - 630. . Prevalence of human papillomavirus infection in the urinary tract of men with urethritis. Int J Urol 17: 563–568, 2010. doi:10.1111/j.1442-2042.2010.02521.x.
Crossref | PubMed | Web of Science | Google Scholar - 631. . Tissue-resident memory T cells. Immunol Rev 255: 165–181, 2013. doi:10.1111/imr.12087.
Crossref | PubMed | Web of Science | Google Scholar - 632. . Localization of porcine reproductive and respiratory syndrome virus infection in boars by in situ riboprobe hybridization. J Vet Sci 3: 87–96, 2002. doi:10.4142/jvs.2002.3.2.87.
Crossref | PubMed | Google Scholar - 633. . Epithelial basal cells are distinct from dendritic cells and macrophages in the mouse epididymis. Biol Reprod 90: 90, 2014. doi:10.1095/biolreprod.113.116681.
Crossref | PubMed | Web of Science | Google Scholar - 634. . Zika virus infects human sertoli cells and modulates the integrity of the in vitro blood-testis barrier model. J Virol 91:
e00623-17 , 2017. doi:10.1128/JVI.00623-17.
Crossref | PubMed | Web of Science | Google Scholar - 635. . Viral infections during pregnancy. Am J Reprod Immunol 73: 199–213, 2015. doi:10.1111/aji.12355.
Crossref | PubMed | Web of Science | Google Scholar - 636. . Hepatitis C-like viruses are produced in cells from rabbit and hare DNA. Sci Rep 5: 14535, 2015. doi:10.1038/srep14535.
Crossref | PubMed | Web of Science | Google Scholar - 637. . A dense network of dendritic cells populates the murine epididymis. Reproduction 141: 653–663, 2011. doi:10.1530/REP-10-0493.
Crossref | PubMed | Web of Science | Google Scholar - 638. . Human retroviral sequences on the Y chromosome. Mol Cell Biol 7: 1559–1562, 1987. doi:10.1128/MCB.7.4.1559.
Crossref | PubMed | Web of Science | Google Scholar - 639. . Human endogenous retrovirus K14C drove genomic diversification of the Y chromosome during primate evolution. J Hum Genet 55: 717–725, 2010. doi:10.1038/jhg.2010.94.
Crossref | PubMed | Web of Science | Google Scholar - 640. . Persistence and clearance of Ebola virus RNA from seminal fluid of Ebola virus disease survivors: a longitudinal analysis and modelling study. Lancet Glob Health 5: e80–e88, 2017. doi:10.1016/S2214-109X(16)30243-1.
Crossref | PubMed | Web of Science | Google Scholar - 641. . Male reproductive disorders and fertility trends: influences of environment and genetic susceptibility. Physiol Rev 96: 55–97, 2016. doi:10.1152/physrev.00017.2015.
Link | Web of Science | Google Scholar - 642. . Detection of DNA viruses in prostate cancer. Sci Rep 6: 25235, 2016. doi:10.1038/srep25235.
Crossref | PubMed | Web of Science | Google Scholar - 643. . Marburg-virus disease in Kenya. Lancet 319: 816–820, 1982. doi:10.1016/S0140-6736(82)91871-2.
Crossref | PubMed | Google Scholar - 644. . The prostate as a reservoir for HIV-1. AIDS 18: 1600–1602, 2004. doi:10.1097/01.aids.0000131364.60081.01.
Crossref | PubMed | Google Scholar - 645. . West Nile virus encephalitis with myositis and orchitis. Hum Pathol 35: 254–258, 2004. doi:10.1016/j.humpath.2003.09.007.
Crossref | PubMed | Web of Science | Google Scholar - 646. . A longitudinal study of Ebola sequelae in Liberia. N Engl J Med 380: 924–934, 2019. doi:10.1056/NEJMoa1805435.
Crossref | PubMed | Web of Science | Google Scholar - 647. . Effect of HSV-2 serostatus on acquisition of HIV by young men: results of a longitudinal study in Orange Farm, South Africa. J Infect Dis 199: 958–964, 2009. doi:10.1086/597208.
Crossref | PubMed | Web of Science | Google Scholar - 648. . Involvement of human endogenous retroviral syncytin-1 in human osteoclast fusion. Bone 48: 837–846, 2011. doi:10.1016/j.bone.2010.11.011.
Crossref | PubMed | Web of Science | Google Scholar - 649. . Prevention of sexual transmission of Ebola in Liberia through a national semen testing and counselling programme for survivors: an analysis of Ebola virus RNA results and behavioural data. Lancet Glob Health 4: e736–e743, 2016. doi:10.1016/S2214-109X(16)30175-9.
Crossref | PubMed | Web of Science | Google Scholar - 650. . Semen alterations in porcine rubulavirus-infected boars are related to viral excretion and have implications for artificial insemination. Res Vet Sci 83: 403–409, 2007. doi:10.1016/j.rvsc.2007.01.018.
Crossref | PubMed | Web of Science | Google Scholar - 651. . Disparate associations of HLA class I markers with HIV-1 acquisition and control of viremia in an African population. PLoS One 6:
e23469 , 2011. doi:10.1371/journal.pone.0023469.
Crossref | PubMed | Web of Science | Google Scholar - 652. . Molecular and structural changes related to hepatitis E virus antigen and its expression in testis inducing apoptosis in Mongolian gerbil model. J Viral Hepat 24: 696–707, 2017. doi:10.1111/jvh.12690.
Crossref | PubMed | Web of Science | Google Scholar - 653. . Sperm-mediated transgenerational inheritance. Front Microbiol 8: 2401, 2017. doi:10.3389/fmicb.2017.02401.
Crossref | PubMed | Web of Science | Google Scholar - 654. . Herpesvirus entry: an update. J Virol 77: 10179–10185, 2003. doi:10.1128/JVI.77.19.10179-10185.2003.
Crossref | PubMed | Web of Science | Google Scholar - 655. . Epithelial cell responses to infection with human papillomavirus. Clin Microbiol Rev 25: 215–222, 2012. doi:10.1128/CMR.05028-11.
Crossref | PubMed | Web of Science | Google Scholar - 656. . Toll-like receptor 3 activation induces antiviral immune responses in mouse sertoli cells. Biol Reprod 79: 766–775, 2008. doi:10.1095/biolreprod.108.068619.
Crossref | PubMed | Web of Science | Google Scholar - 657. . Zika virus in the male reproductive tract. Viruses 10: 198, 2018. doi:10.3390/v10040198.
Crossref | PubMed | Web of Science | Google Scholar - 658. . Digital expression profiles of human endogenous retroviral families in normal and cancerous tissues. Cancer Immun 4: 2, 2004.
PubMed | Google Scholar - 659. . Mucin 6 in seminal plasma binds DC-SIGN and potently blocks dendritic cell mediated transfer of HIV-1 to CD4(+) T-lymphocytes. Virology 391: 203–211, 2009. doi:10.1016/j.virol.2009.06.011.
Crossref | PubMed | Web of Science | Google Scholar - 660. . Phenotype and function of peripheral and prostatic lymphocytes in patients with benign prostatic hyperplasia. J Urol 151: 480–484, 1994. doi:10.1016/S0022-5347(17)34998-4.
Crossref | PubMed | Web of Science | Google Scholar - 661. . Genital infections due to herpes simplex virus type 2 in male guinea pigs. J Infect Dis 159: 89–95, 1989. doi:10.1093/infdis/159.1.89.
Crossref | PubMed | Web of Science | Google Scholar - 662. . Th17 cells are preferentially infected very early after vaginal transmission of SIV in macaques. Cell Host Microbe 19: 529–540, 2016. doi:10.1016/j.chom.2016.03.005.
Crossref | PubMed | Web of Science | Google Scholar - 663. . Transcription of hepatitis B virus in peripheral blood mononuclear cells from persistently infected patients. J Virol 71: 5399–5407, 1997. doi:10.1128/JVI.71.7.5399-5407.1997.
Crossref | PubMed | Web of Science | Google Scholar - 664. . Asymptomatic virus shedding in men with genital herpes infection. Scand J Infect Dis 18: 195–197, 1986. doi:10.3109/00365548609032327.
Crossref | PubMed | Google Scholar - 665. . Immunoprofiles of human Sertoli cells infected with Zika virus reveals unique insights into host-pathogen crosstalk. Sci Rep 8: 8702, 2018. doi:10.1038/s41598-018-27027-7.
Crossref | PubMed | Web of Science | Google Scholar - 666. . Hepatitis B virus infection and the risk of male infertility: a population-based analysis. Fertil Steril 102: 1677–1684, 2014. doi:10.1016/j.fertnstert.2014.09.017.
Crossref | PubMed | Web of Science | Google Scholar - 667. . Sertoli cell-initiated testicular innate immune response through Toll-like receptor-3 activation is negatively regulated by Tyro3, Axl, and mer receptors. Endocrinology 151: 2886–2897, 2010. doi:10.1210/en.2009-1498.
Crossref | PubMed | Web of Science | Google Scholar - 668. . Spinal neurons involved in the control of the seminal vesicles: a transsynaptic labeling study using pseudorabies virus in rats. Neuroscience 158: 786–797, 2009. doi:10.1016/j.neuroscience.2008.10.008.
Crossref | PubMed | Web of Science | Google Scholar - 669. . Human papillomavirus infection of the epididymis and ductus deferens an evaluation by nested polymerase chain reaction. Arch Pathol Lab Med 127: 1471–1474, 2003.
PubMed | Web of Science | Google Scholar - 670. . Studies on the propagation in vitro of poliomyelitis viruses. I. Viral multiplications in tissue cultures employing monkey and human testicular cells. J Exp Med 96: 355–367, 1952. doi:10.1084/jem.96.4.355.
Crossref | PubMed | Web of Science | Google Scholar - 671. . Detection and quantification of HIV-1 in semen: identification of a subpopulation of men at high potential risk of viral sexual transmission. AIDS 13: 823–831, 1999. doi:10.1097/00002030-199905070-00012.
Crossref | PubMed | Web of Science | Google Scholar - 672. . Human herpesvirus 6 (HHV-6) is transmitted from parent to child in an integrated form and characterization of cases with chromosomally integrated HHV-6 DNA. J Med Virol 73: 465–473, 2004. doi:10.1002/jmv.20113.
Crossref | PubMed | Web of Science | Google Scholar - 673. . Seminal plasma and semen amyloids enhance cytomegalovirus infection in cell culture. J Virol 87: 12583–12591, 2013. doi:10.1128/JVI.02083-13.
Crossref | PubMed | Web of Science | Google Scholar - 674. . Retroviral invasion of the koala genome. Nature 442: 79–81, 2006. doi:10.1038/nature04841.
Crossref | PubMed | Web of Science | Google Scholar - 675. . Prostate cancer and sexually transmitted diseases: a meta-analysis. Fam Med 37: 506–512, 2005.
PubMed | Web of Science | Google Scholar - 676. . Herpes simplex virus. Front Biosci 7: d752–d764, 2002. doi:10.2741/taylor.
Crossref | PubMed | Google Scholar - 677. . Replication of equid herpesvirus-1 (EHV-1) in the testes and epididymides of ponies and venereal shedding of infectious virus. J Comp Pathol 115: 385–397, 1996. doi:10.1016/S0021-9975(96)80073-9.
Crossref | PubMed | Web of Science | Google Scholar - 678. . Superfibronectin, a multimeric form of fibronectin, increases HIV infection of primary CD4+ T lymphocytes. J Immunol 164: 3236–3245, 2000. doi:10.4049/jimmunol.164.6.3236.
Crossref | PubMed | Web of Science | Google Scholar - 679. . A congenital persistent infection of bovine virus diarrhoea virus in pigs: clinical, virological and immunological observations. Vet Q 19: 97–101, 1997. doi:10.1080/01652176.1997.9694750.
Crossref | PubMed | Web of Science | Google Scholar - 680. . Swine diseases transmissible with artificial insemination. J Am Vet Med Assoc 185: 511–516, 1984.
PubMed | Web of Science | Google Scholar - 681. . Tumour necrosis factor-alpha released by testicular macrophages induces apoptosis of germ cells in autoimmune orchitis. Hum Reprod 23: 1865–1872, 2008. doi:10.1093/humrep/den240.
Crossref | PubMed | Web of Science | Google Scholar - 682. . Spontaneous encephalomyelitis of mice, a new virus disease. J Exp Med 65: 705–719, 1937. doi:10.1084/jem.65.5.705.
Crossref | PubMed | Google Scholar - 683. . Phenotypic characterization of infiltrating leukocytes in benign prostatic hyperplasia. Lab Invest 66: 96–107, 1992.
PubMed | Web of Science | Google Scholar - 684. . Demonstration of the carrier state in naturally acquired equine arteritis virus infection in the stallion. Res Vet Sci 41: 279–280, 1986. doi:10.1016/S0034-5288(18)30616-7.
Crossref | PubMed | Web of Science | Google Scholar - 685. . Herpes simplex virus type 2 and syphilis infections with HIV: an evolving synergy in transmission and prevention. Curr Opin HIV AIDS 4: 294–299, 2009. doi:10.1097/COH.0b013e32832c1881.
Crossref | PubMed | Web of Science | Google Scholar - 686. . Human papillomavirus incidence and clearance among HIV-positive and HIV-negative men in sub-Saharan Africa. AIDS 26: 1555–1565, 2012. doi:10.1097/QAD.0b013e328353b83c.
Crossref | PubMed | Web of Science | Google Scholar - 687. . Male circumcision for the prevention of HSV-2 and HPV infections and syphilis. N Engl J Med 360: 1298–1309, 2009. doi:10.1056/NEJMoa0802556.
Crossref | PubMed | Web of Science | Google Scholar - 688. . Characterization of lymphocytes in the adult rat testis by flow cytometry: effects of activin and transforming growth factor beta on lymphocyte subsets in vitro. Biol Reprod 58: 943–951, 1998. doi:10.1095/biolreprod58.4.943.
Crossref | PubMed | Web of Science | Google Scholar - 689. . Tight junctions go viral! Viruses 7: 5145–5154, 2015. doi:10.3390/v7092865.
Crossref | PubMed | Web of Science | Google Scholar - 690. . Enhancement of herpes simplex virus (HSV) infection by seminal plasma and semen amyloids implicates a new target for the prevention of HSV infection. Viruses 7: 2057–2073, 2015. doi:10.3390/v7042057.
Crossref | PubMed | Web of Science | Google Scholar - 691. . HIV infection of the male genital tract--consequences for sexual transmission and reproduction. Int J Androl 33: e98–e108, 2010. doi:10.1111/j.1365-2605.2009.00973.x.
Crossref | PubMed | Google Scholar - 692. . Antiviral responses of human Leydig cells to mumps virus infection or poly I:C stimulation. Hum Reprod 23: 2095–2103, 2008. doi:10.1093/humrep/den207.
Crossref | PubMed | Web of Science | Google Scholar - 693. . Infection of semen-producing organs by SIV during the acute and chronic stages of the disease. PLoS One 3:
e1792 , 2008. doi:10.1371/journal.pone.0001792.
Crossref | PubMed | Web of Science | Google Scholar - 694. . Human prostate supports more efficient replication of HIV-1 R5 than X4 strains ex vivo. Retrovirology 5: 119, 2008. doi:10.1186/1742-4690-5-119.
Crossref | PubMed | Web of Science | Google Scholar - 695. . The male as a reservoir of infection with cytomegalovirus, herpes and mycoplasma. N Engl J Med 289: 697–698, 1973. doi:10.1056/NEJM197309272891321.
Crossref | PubMed | Web of Science | Google Scholar - 696. . Routes of Zika virus dissemination in the testis and epididymis of immunodeficient mice. Nat Commun 9: 5350, 2018. doi:10.1038/s41467-018-07782-x.
Crossref | PubMed | Web of Science | Google Scholar - 697. . Dose-dependent effects of recombinant human interleukin-6 on the pituitary-testicular axis. J Interferon Cytokine Res 19: 1271–1276, 1999. doi:10.1089/107999099312948.
Crossref | PubMed | Web of Science | Google Scholar - 698. . Egress of sperm autoantigen from seminiferous tubules maintains systemic tolerance. J Clin Invest 127: 1046–1060, 2017. doi:10.1172/JCI89927.
Crossref | PubMed | Web of Science | Google Scholar - 699. . Distribution of caprine arthritis encephalitis virus provirus, RNA, and antigen in the reproductive tract of one naturally and seven experimentally infected bucks. Theriogenology 80: 933–939, 2013. doi:10.1016/j.theriogenology.2013.07.021.
Crossref | PubMed | Web of Science | Google Scholar - 700. . Shedding of hepatitis C virus in semen of human immunodeficiency virus-infected men. Open Forum Infect Dis 3:
ofw057 , 2016. doi:10.1093/ofid/ofw057.
Crossref | PubMed | Web of Science | Google Scholar - 701. . Global HIV & AIDS statistics — 2018 fact sheet [Online]. 2019. https://www.unaids.org/en/resources/fact-sheet.
Google Scholar - 702. . Zika virus causes testicular atrophy. Sci Adv 3:
e1602899 , 2017. doi:10.1126/sciadv.1602899.
Crossref | PubMed | Web of Science | Google Scholar - 703. . Fetal growth restriction caused by sexual transmission of Zika virus in mice. J Infect Dis 215: 1720–1724, 2017. doi:10.1093/infdis/jix204.
Crossref | PubMed | Web of Science | Google Scholar - 704. . Ebola virus persistence in semen of male survivors. Clin Infect Dis 62: 1552–1555, 2016. doi:10.1093/cid/ciw202.
Crossref | PubMed | Web of Science | Google Scholar - 705. . Expression of toll-like receptor-9 is increased in poorly differentiated prostate tumors. Prostate 70: 817–824, 2010. doi:10.1002/pros.21115.
Crossref | PubMed | Web of Science | Google Scholar - 706. . A duplex real-time RT-PCR for the detection of bluetongue virus in bovine semen. J Virol Methods 169: 162–168, 2010. doi:10.1016/j.jviromet.2010.07.019.
Crossref | PubMed | Web of Science | Google Scholar - 707. . Distinct cytokine/chemokine network in semen and blood characterize different stages of HIV infection. AIDS 30: 193–201, 2016. doi:10.1097/QAD.0000000000000964.
Crossref | PubMed | Web of Science | Google Scholar - 708. . HIV is not transmitted under fully suppressive therapy: The Swiss Statement--eight years later. Swiss Med Wkly 146:
w14246 , 2016. doi:10.4414/smw.2016.14246.
Crossref | PubMed | Web of Science | Google Scholar - 709. . Evaluation of cystatin C activities against HIV. Indian J Med Res 141: 423–430, 2015. doi:10.4103/0971-5916.159282.
Crossref | PubMed | Web of Science | Google Scholar - 710. . Cell culture models for the investigation of hepatitis B and D virus infection. Viruses 8: 261, 2016. doi:10.3390/v8090261.
Crossref | PubMed | Web of Science | Google Scholar - 711. . Ebola virus shedding and transmission: review of current evidence. J Infect Dis 214, Suppl 3: S177–S184, 2016. doi:10.1093/infdis/jiw254.
Crossref | PubMed | Web of Science | Google Scholar - 712. . Sperm DNA damage in patients with chronic viral C hepatitis. Eur J Intern Med 23: e19–e24, 2012. doi:10.1016/j.ejim.2011.08.011.
Crossref | PubMed | Web of Science | Google Scholar - 713. . Male accessory gland infection and sperm parameters (review). Int J Androl 34: e330–e347, 2011. doi:10.1111/j.1365-2605.2011.01200.x.
Crossref | PubMed | Google Scholar - 714. . The potential for sexual transmission to compromise control of Ebola virus outbreaks. Biol Lett 12:
20151079 , 2016. doi:10.1098/rsbl.2015.1079.
Crossref | PubMed | Web of Science | Google Scholar - 715. . Infection and replication of herpes simplex virus type 1 in an organotypic epithelial culture system. Virology 230: 236–243, 1997. doi:10.1006/viro.1997.8484.
Crossref | PubMed | Web of Science | Google Scholar - 716. . Persistent bovine pestivirus infection localized in the testes of an immuno-competent, non-viraemic bull. Vet Microbiol 61: 165–175, 1998. doi:10.1016/S0378-1135(98)00177-1.
Crossref | PubMed | Web of Science | Google Scholar - 717. . The enemy within: endogenous retroelements and autoimmune disease. Nat Immunol 15: 415–422, 2014. doi:10.1038/ni.2872.
Crossref | PubMed | Web of Science | Google Scholar - 718. . Emerging zoonotic viral infections of occupational health importance. Pathog Dis 77:
ftz018 , 2019. doi:10.1093/femspd/ftz018.
Crossref | PubMed | Web of Science | Google Scholar - 719. . Epididymitis caused by coxsackievirus A6 in association with hand, foot, and mouth disease. J Clin Microbiol 52: 4412–4413, 2014. doi:10.1128/JCM.02441-14.
Crossref | PubMed | Web of Science | Google Scholar - 720. . Hepatitis B virus incidence and risk factors among human immunodeficiency virus-1 negative men who have sex with men in Kenya. Open Forum Infect Dis 4:
ofw253 , 2016. doi:10.1093/ofid/ofw253.
Crossref | PubMed | Web of Science | Google Scholar - 721. . Venereal shedding of equid herpesvirus-1 (EHV-1) in naturally infected stallions. J Vet Intern Med 26: 1500–1504, 2012. doi:10.1111/j.1939-1676.2012.00997.x.
Crossref | PubMed | Web of Science | Google Scholar - 722. . The integrated HIV-1 provirus in patient sperm chromosome and its transfer into the early embryo by fertilization. PLoS One 6:
e28586 , 2011. doi:10.1371/journal.pone.0028586.
Crossref | PubMed | Web of Science | Google Scholar - 723. . The role of dendritic cells in male reproductive tract. Am J Reprod Immunol 76: 186–192, 2016. doi:10.1111/aji.12536.
Crossref | PubMed | Web of Science | Google Scholar - 724. . Toll-like receptor 3-initiated antiviral responses in mouse male germ cells in vitro. Biol Reprod 86: 106, 2012. doi:10.1095/biolreprod.111.096719.
Crossref | PubMed | Web of Science | Google Scholar - 725. . Human herpesvirus 6 chromosomal integration in immunocompetent patients results in high levels of viral DNA in blood, sera, and hair follicles. J Clin Microbiol 44: 1571–1574, 2006. doi:10.1128/JCM.44.4.1571-1574.2006.
Crossref | PubMed | Web of Science | Google Scholar - 726. . In vivo genetic manipulation of spermatogonial stem cells and their microenvironment by adeno-associated viruses. Stem Cell Reports 10: 1551–1564, 2018. doi:10.1016/j.stemcr.2018.03.005.
Crossref | PubMed | Web of Science | Google Scholar - 727. . Environmental epigenetic inheritance through gametes and implications for human reproduction. Hum Reprod Update 21: 194–208, 2015. doi:10.1093/humupd/dmu061.
Crossref | PubMed | Web of Science | Google Scholar - 728. . Male urogenital infections: impact of infection and inflammation on ejaculate parameters. World J Urol 31: 717–723, 2013. doi:10.1007/s00345-013-1082-7.
Crossref | PubMed | Web of Science | Google Scholar - 729. . Human papillomavirus and Epstein Barr virus in prostate cancer: Koilocytes indicate potential oncogenic influences of human papillomavirus in prostate cancer. Prostate 73: 236–241, 2013. doi:10.1002/pros.22562.
Crossref | PubMed | Web of Science | Google Scholar - 730. . Active Ebola virus replication and heterogeneous evolutionary rates in EVD survivors. Cell Rep 22: 1159–1168, 2018. doi:10.1016/j.celrep.2018.01.008.
Crossref | PubMed | Web of Science | Google Scholar - 731. . WHO | New data on male circumcision and HIV prevention:Research implications for policy and programming. In: WHO/UNAIDS. Montreux: World Health Organization, 2007.
Google Scholar - 732. . Hepatitis B [Online]. 2018. https://www.who.int/news-room/fact-sheets/detail/hepatitis-b.
Google Scholar - 733. . Hepatitis D [Online]. 2018. https://www.who.int/news-room/fact-sheets/detail/hepatitis-d.
Google Scholar - 734. . Herpes simplex virus [Online]. 2019. https://www.who.int/news-room/fact-sheets/detail/herpes-simplex-virus.
Google Scholar - 735. . HIV/AIDS Key Facts [Online]. 2019. https://www.who.int/news-room/fact-sheets/detail/hiv-aids.
Google Scholar - 736. . Localization of lactoferrin in the male reproductive tract. Int J Androl 12: 179–186, 1989. doi:10.1111/j.1365-2605.1989.tb01302.x.
Crossref | PubMed | Google Scholar - 737. . Discovery of unfixed endogenous retrovirus insertions in diverse human populations. Proc Natl Acad Sci USA 113: E2326–E2334, 2016. doi:10.1073/pnas.1602336113.
Crossref | PubMed | Web of Science | Google Scholar - 738. . Human Leydig cells are productively infected by some HIV-2 and SIV strains but not by HIV-1. AIDS 17: 183–188, 2003. doi:10.1097/00002030-200301240-00007.
Crossref | PubMed | Web of Science | Google Scholar - 739. . Phenotypic and functional heterogeneity of the testicular macrophage population: a new regulatory model. J Reprod Immunol 97: 147–158, 2013. doi:10.1016/j.jri.2013.01.001.
Crossref | PubMed | Web of Science | Google Scholar - 740. . Simian immunodeficiency virus infection and immune responses in the pig-tailed macaque testis. J Leukoc Biol 97: 599–609, 2015. doi:10.1189/jlb.4A0914-438R.
Crossref | PubMed | Web of Science | Google Scholar - 741. . Unusually high HIV infectiousness in an HIV-, HCV- and HSV-2-coinfected heterosexual man. Swiss Med Wkly 139: 207–209, 2009.
PubMed | Web of Science | Google Scholar - 742. . Tissue reservoirs of HIV. Curr Opin HIV AIDS 11: 362–370, 2016. doi:10.1097/COH.0000000000000293.
Crossref | PubMed | Web of Science | Google Scholar - 743. . Human viruses: discovery and emergence. Philos Trans R Soc Lond B Biol Sci 367: 2864–2871, 2012. doi:10.1098/rstb.2011.0354.
Crossref | PubMed | Web of Science | Google Scholar - 744. . WHO fact sheets Herpes simplex virus [Online]. 2017. https://www.who.int/news-room/fact-sheets/detail/herpes-simplex-virus.
Google Scholar - 745. . Global AIDS Update 2016. UNAIDS Rep 422, 2016.
Google Scholar - 746. . Mumps virus-induced innate immune responses in mouse Sertoli and Leydig cells. Sci Rep 6: 19507, 2016. doi:10.1038/srep19507.
Crossref | PubMed | Web of Science | Google Scholar - 747. . Expression patterns and functions of Toll-like receptors in mouse sertoli cells. Endocrinology 149: 4402–4412, 2008. doi:10.1210/en.2007-1776.
Crossref | PubMed | Web of Science | Google Scholar - 748. . Mouse testicular cell type-specific antiviral response against mumps virus replication. Front Immunol 8: 117, 2017. doi:10.3389/fimmu.2017.00117.
Crossref | PubMed | Web of Science | Google Scholar - 749. . Atypical microorganisms in expressed prostatic secretion from patients with chronic prostatitis/chronic pelvic pain syndrome: microbiological results from a case-control study. Urol Int 91: 410–416, 2013. doi:10.1159/000350934.
Crossref | PubMed | Web of Science | Google Scholar - 750. . Role of IFI 16, a member of the interferon-inducible p200-protein family, in prostate epithelial cellular senescence. Oncogene 22: 4831–4840, 2003. doi:10.1038/sj.onc.1206754.
Crossref | PubMed | Web of Science | Google Scholar - 751. . Gas6 and the Tyro 3 receptor tyrosine kinase subfamily regulate the phagocytic function of Sertoli cells. Reproduction 135: 77–87, 2008. doi:10.1530/REP-07-0287.
Crossref | PubMed | Web of Science | Google Scholar - 752. . The role of tumor necrosis factor-alpha in the regulation of mouse Leydig cell steroidogenesis. Endocrinology 132: 2438–2444, 1993. doi:10.1210/endo.132.6.8504748.
Crossref | PubMed | Web of Science | Google Scholar - 753. . Trends in herpes simplex virus type 1 and type 2 seroprevalence in the United States. JAMA 296: 964–973, 2006. doi:10.1001/jama.296.8.964.
Crossref | PubMed | Web of Science | Google Scholar - 754. . Endogenous retroviruses of non-avian/mammalian vertebrates illuminate diversity and deep history of retroviruses. PLoS Pathog 14:
e1007072 , 2018. doi:10.1371/journal.ppat.1007072.
Crossref | PubMed | Web of Science | Google Scholar - 755. . Identification of multiple novel epididymis-specific beta-defensin isoforms in humans and mice. J Immunol 169: 2516–2523, 2002. doi:10.4049/jimmunol.169.5.2516.
Crossref | PubMed | Web of Science | Google Scholar - 756. . Worldwide prevalence of human papillomavirus and relative risk of prostate cancer: a meta-analysis. Sci Rep 5: 14667, 2015. doi:10.1038/srep14667.
Crossref | PubMed | Web of Science | Google Scholar - 757. . Effect of hepatitis C virus infection on the outcomes of in vitro fertilization. Int J Clin Exp Med 8: 6230–6235, 2015.
PubMed | Web of Science | Google Scholar - 758. . Cytomegalovirus infection and viral shedding in the genital tract of infertile couples. J Med Virol 45: 179–182, 1995. doi:10.1002/jmv.1890450212.
Crossref | PubMed | Web of Science | Google Scholar - 759. . The TRIM5 gene modulates penile mucosal acquisition of simian immunodeficiency virus in rhesus monkeys. J Virol 85: 10389–10398, 2011. doi:10.1128/JVI.00854-11.
Crossref | PubMed | Web of Science | Google Scholar - 760. . Association between human papillomavirus and prostate cancer: A meta-analysis. Oncol Lett 14: 1855–1865, 2017. doi:10.3892/ol.2017.6367.
Crossref | PubMed | Web of Science | Google Scholar - 761. . Cytomegalovirus prostatitis: a series of 4 cases. Int J Surg Pathol 18: 55–59, 2010. doi:10.1177/1066896908321182.
Crossref | PubMed | Web of Science | Google Scholar - 762. . Testicular cancer and viral infections: a systematic literature review and meta-analysis. J Med Virol 85: 2165–2175, 2013. doi:10.1002/jmv.23704.
Crossref | PubMed | Web of Science | Google Scholar - 763. . Effectiveness of semen washing to prevent human immunodeficiency virus (HIV) transmission and assist pregnancy in HIV-discordant couples: a systematic review and meta-analysis. Fertil Steril 105: 645–655.e2, 2016. doi:10.1016/j.fertnstert.2015.11.028.
Crossref | PubMed | Web of Science | Google Scholar - 764. . Detection of human polyomaviruses and papillomaviruses in prostatic tissue reveals the prostate as a habitat for multiple viral infections. Prostate 53: 263–276, 2002. doi:10.1002/pros.10157.
Crossref | PubMed | Web of Science | Google Scholar - 765. . The International Committee for Monitoring Assisted Reproductive Technology (ICMART) and the World Health Organization (WHO) revised glossary on ART terminology, 2009. Hum Reprod 24: 2683–2687, 2009. doi:10.1093/humrep/dep343.
Crossref | PubMed | Web of Science | Google Scholar - 766. . Identification and pathological characterization of persistent asymptomatic Ebola virus infection in rhesus monkeys. Nat Microbiol 2: 17113, 2017. doi:10.1038/nmicrobiol.2017.113.
Crossref | PubMed | Web of Science | Google Scholar - 767. . Central nervous system neurons labeled following the injection of pseudorabies virus into the rat prostate gland. Prostate 44: 240–247, 2000. doi:10.1002/1097-0045(20000801)44:3<240:AID-PROS9>3.0.CO;2-9.
Crossref | PubMed | Web of Science | Google Scholar - 768. . CRISPR/Cas9-mediated genome editing reveals the synergistic effects of β-defensin family members on sperm maturation in rat epididymis. FASEB J 32: 1354–1363, 2018. doi:10.1096/fj.201700936R.
Crossref | PubMed | Web of Science | Google Scholar - 769. . Inherited chromosomally integrated human herpesvirus 6 genomes are ancient, intact, and potentially able to reactivate from telomeres. J Virol 91:
01137-17 , 2017. doi:10.1128/JVI.01137-17.
Crossref | PubMed | Web of Science | Google Scholar - 770. . [Detection of severe acute respiratory syndrome (SARS)-associated coronavirus RNA in autopsy tissues with in situ hybridization]. J First Mil Med Univ 23: 1125–1127, 2003.
Google Scholar - 771. . Testicular defense systems: immune privilege and innate immunity. Cell Mol Immunol 11: 428–437, 2014. doi:10.1038/cmi.2014.38.
Crossref | PubMed | Web of Science | Google Scholar - 772. . Integration of viral genomes. Nature 256: 471–473, 1975. doi:10.1038/256471a0.
Crossref | PubMed | Web of Science | Google Scholar - 773. . [Integration and transfection of an arbovirus by mammalian cells]. Mol Biol (Mosk) 10: 1296–1302, 1976.
PubMed | Web of Science | Google Scholar - 774. . An epididymis-specific β-defensin is important for the initiation of sperm maturation. Nat Cell Biol 6: 458–464, 2004. doi:10.1038/ncb1127.
Crossref | PubMed | Web of Science | Google Scholar - 775. . Epigallocatechin gallate inhibits Macaque SEVI-mediated enhancement of SIV or SHIV infection. J Acquir Immune Defic Syndr 75: 232–240, 2017. doi:10.1097/QAI.0000000000001361.
Crossref | PubMed | Web of Science | Google Scholar - 776. . HIV-1 efficient entry in inner foreskin is mediated by elevated CCL5/RANTES that recruits T cells and fuels conjugate formation with Langerhans cells. PLoS Pathog 7:
e1002100 , 2011. doi:10.1371/journal.ppat.1002100.
Crossref | PubMed | Web of Science | Google Scholar - 777. . The HIV-1 viral synapse signals human foreskin keratinocytes to secrete thymic stromal lymphopoietin facilitating HIV-1 foreskin entry. Mucosal Immunol 11: 158–171, 2018. doi:10.1038/mi.2017.23.
Crossref | PubMed | Web of Science | Google Scholar - 778. . Immune surveillance by CD8αα+ skin-resident T cells in human herpes virus infection. [Correction in Nature 500: 242, 2013.] Nature 497: 494–497, 2013. doi:10.1038/nature12110.
Crossref | PubMed | Web of Science | Google Scholar - 779. . RIG-I-like receptors mediate innate antiviral response in mouse testis. Mol Endocrinol 27: 1455–1467, 2013. doi:10.1210/me.2013-1075.
Crossref | PubMed | Google Scholar - 780. . p204-initiated innate antiviral response in mouse Leydig cells. Biol Reprod 91: 8, 2014. doi:10.1095/biolreprod.114.119396.
Crossref | PubMed | Web of Science | Google Scholar - 781. . Pattern recognition receptor-initiated innate antiviral responses in mouse epididymal epithelial cells. J Immunol 194: 4825–4835, 2015. doi:10.4049/jimmunol.1402706.
Crossref | PubMed | Web of Science | Google Scholar - 782. . Relationship between circumcision and human papillomavirus infection: a systematic review and meta-analysis. Asian J Androl 19: 125–131, 2017. doi:10.4103/1008-682X.175092.
Crossref | PubMed | Web of Science | Google Scholar - 783. . Integration of foreign DNA sequences into mouse sperm genome. DNA Cell Biol 16: 291–300, 1997. doi:10.1089/dna.1997.16.291.
Crossref | PubMed | Web of Science | Google Scholar