Antiviral drug

Last updated

Antiretroviral drugs for HIV Antiretroviral Drugs to Treat HIV Infection (31793869534).jpg
Antiretroviral drugs for HIV

Antiviral drugs are a class of medication used for treating viral infections. [1] Most antivirals target specific viruses, while a broad-spectrum antiviral is effective against a wide range of viruses. [2] Antiviral drugs are a class of antimicrobials, a larger group which also includes antibiotic (also termed antibacterial), antifungal and antiparasitic drugs, [3] or antiviral drugs based on monoclonal antibodies. [4] Most antivirals are considered relatively harmless to the host, and therefore can be used to treat infections. They should be distinguished from virucides, which are not medication but deactivate or destroy virus particles, either inside or outside the body. Natural virucides are produced by some plants such as eucalyptus and Australian tea trees. [5]

Contents

Medical uses

Most of the antiviral drugs now available are designed to help deal with HIV, herpes viruses, the hepatitis B and C viruses, and influenza A and B viruses. [6]

Viruses use the host's cells to replicate and this makes it difficult to find targets for the drug that would interfere with the virus without also harming the host organism's cells. Moreover, the major difficulty in developing vaccines and antiviral drugs is due to viral variation. [7]

The emergence of antivirals is the product of a greatly expanded knowledge of the genetic and molecular function of organisms, allowing biomedical researchers to understand the structure and function of viruses, major advances in the techniques for finding new drugs, and the pressure placed on the medical profession to deal with the human immunodeficiency virus (HIV), the cause of acquired immunodeficiency syndrome (AIDS).[ citation needed ]

The first experimental antivirals were developed in the 1960s, mostly to deal with herpes viruses, and were found using traditional trial-and-error drug discovery methods. Researchers grew cultures of cells and infected them with the target virus. They then introduced into the cultures chemicals which they thought might inhibit viral activity and observed whether the level of virus in the cultures rose or fell. Chemicals that seemed to have an effect were selected for closer study.[ citation needed ]

This was a very time-consuming, hit-or-miss procedure, and in the absence of a good knowledge of how the target virus worked, it was not efficient in discovering effective antivirals which had few side effects. Only in the 1980s, when the full genetic sequences of viruses began to be unraveled, did researchers begin to learn how viruses worked in detail, and exactly what chemicals were needed to thwart their reproductive cycle. [8]

Antiviral drug design

Antiviral targeting

The general idea behind modern antiviral drug design is to identify viral proteins, or parts of proteins, that can be disabled. These "targets" should generally be as unlike any proteins or parts of proteins in humans as possible, to reduce the likelihood of side effects. The targets should also be common across many strains of a virus, or even among different species of virus in the same family, so a single drug will have broad effectiveness. For example, a researcher might target a critical enzyme synthesized by the virus, but not by the patient, that is common across strains, and see what can be done to interfere with its operation.

Once targets are identified, candidate drugs can be selected, either from drugs already known to have appropriate effects or by actually designing the candidate at the molecular level with a computer-aided design program.

The target proteins can be manufactured in the lab for testing with candidate treatments by inserting the gene that synthesizes the target protein into bacteria or other kinds of cells. The cells are then cultured for mass production of the protein, which can then be exposed to various treatment candidates and evaluated with "rapid screening" technologies.

Approaches by virus life cycle stage

Viruses consist of a genome and sometimes a few enzymes stored in a capsule made of protein (called a capsid), and sometimes covered with a lipid layer (sometimes called an 'envelope'). Viruses cannot reproduce on their own and instead propagate by subjugating a host cell to produce copies of themselves, thus producing the next generation.

Researchers working on such "rational drug design" strategies for developing antivirals have tried to attack viruses at every stage of their life cycles. Some species of mushrooms have been found to contain multiple antiviral chemicals with similar synergistic effects. [9] Compounds isolated from fruiting bodies and filtrates of various mushrooms have broad-spectrum antiviral activities, but successful production and availability of such compounds as frontline antiviral is a long way away. [10]

Viral life cycles vary in their precise details depending on the type of virus, but they all share a general pattern:

  1. Attachment to a host cell.
  2. Release of viral genes and possibly enzymes into the host cell.
  3. Replication of viral components using host-cell machinery.
  4. Assembly of viral components into complete viral particles.
  5. Release of viral particles to infect new host cells.

Before cell entry

One antiviral strategy is to interfere with the ability of a virus to infiltrate a target cell. The virus must go through a sequence of steps to do this, beginning with binding to a specific "receptor" molecule on the surface of the host cell and ending with the virus "uncoating" inside the cell and releasing its contents. Viruses that have a lipid envelope must also fuse their envelope with the target cell, or with a vesicle that transports them into the cell before they can uncoat.

This stage of viral replication can be inhibited in two ways:

  1. Using agents which mimic the virus-associated protein (VAP) and bind to the cellular receptors. This may include VAP anti-idiotypic antibodies, natural ligands of the receptor and anti-receptor antibodies.[ clarification needed ]
  2. Using agents which mimic the cellular receptor and bind to the VAP. This includes anti-VAP antibodies, receptor anti-idiotypic antibodies, extraneous receptor and synthetic receptor mimics.

This strategy of designing drugs can be very expensive, and since the process of generating anti-idiotypic antibodies is partly trial and error, it can be a relatively slow process until an adequate molecule is produced.

Entry inhibitor

A very early stage of viral infection is viral entry, when the virus attaches to and enters the host cell. A number of "entry-inhibiting" or "entry-blocking" drugs are being developed to fight HIV. HIV most heavily targets a specific type of lymphocyte known as "helper T cells", and identifies these target cells through T-cell surface receptors designated "CD4" and "CCR5". Attempts to interfere with the binding of HIV with the CD4 receptor have failed to stop HIV from infecting helper T cells, but research continues on trying to interfere with the binding of HIV to the CCR5 receptor in hopes that it will be more effective.

HIV infects a cell through fusion with the cell membrane, which requires two different cellular molecular participants, CD4 and a chemokine receptor (differing depending on the cell type). Approaches to blocking this virus/cell fusion have shown some promise in preventing entry of the virus into a cell. At least one of these entry inhibitors—a biomimetic peptide called Enfuvirtide, or the brand name Fuzeon—has received FDA approval and has been in use for some time. Potentially, one of the benefits from the use of an effective entry-blocking or entry-inhibiting agent is that it potentially may not only prevent the spread of the virus within an infected individual but also the spread from an infected to an uninfected individual.

One possible advantage of the therapeutic approach of blocking viral entry (as opposed to the currently dominant approach of viral enzyme inhibition) is that it may prove more difficult for the virus to develop resistance to this therapy than for the virus to mutate or evolve its enzymatic protocols.

Uncoating inhibitors

Inhibitors of uncoating have also been investigated. [11] [12]

Amantadine and rimantadine have been introduced to combat influenza. These agents act on penetration and uncoating. [13]

Pleconaril works against rhinoviruses, which cause the common cold, by blocking a pocket on the surface of the virus that controls the uncoating process. This pocket is similar in most strains of rhinoviruses and enteroviruses, which can cause diarrhea, meningitis, conjunctivitis, and encephalitis. [14]

Some scientists are making the case that a vaccine against rhinoviruses, the predominant cause of the common cold, is achievable. Vaccines that combine dozens of varieties of rhinovirus at once are effective in stimulating antiviral antibodies in mice and monkeys, researchers reported in Nature Communications in 2016. [15]

Rhinoviruses are the most common cause of the common cold; other viruses such as respiratory syncytial virus, parainfluenza virus and adenoviruses can cause them too. [16] Rhinoviruses also exacerbate asthma attacks. Although rhinoviruses come in many varieties, they do not drift to the same degree that influenza viruses do. A mixture of 50 inactivated rhinovirus types should be able to stimulate neutralizing antibodies against all of them to some degree. [17]

During viral synthesis

A second approach is to target the processes that synthesize virus components after a virus invades a cell.

Reverse transcription

One way of doing this is to develop nucleotide or nucleoside analogues that look like the building blocks of RNA or DNA, but deactivate the enzymes that synthesize the RNA or DNA once the analogue is incorporated. This approach is more commonly associated with the inhibition of reverse transcriptase (RNA to DNA) than with "normal" transcriptase (DNA to RNA).

The first successful antiviral, aciclovir, is a nucleoside analogue, and is effective against herpesvirus infections. The first antiviral drug to be approved for treating HIV, zidovudine (AZT), is also a nucleoside analogue.

An improved knowledge of the action of reverse transcriptase has led to better nucleoside analogues to treat HIV infections. One of these drugs, lamivudine, has been approved to treat hepatitis B, which uses reverse transcriptase as part of its replication process. Researchers have gone further and developed inhibitors that do not look like nucleosides, but can still block reverse transcriptase.

Another target being considered for HIV antivirals include RNase H—which is a component of reverse transcriptase that splits the synthesized DNA from the original viral RNA.

Integrase

Another target is integrase, which integrate the synthesized DNA into the host cell genome. Examples of integrase inhibitors include raltegravir, elvitegravir, and dolutegravir.

Transcription

Once a virus genome becomes operational in a host cell, it then generates messenger RNA (mRNA) molecules that direct the synthesis of viral proteins. Production of mRNA is initiated by proteins known as transcription factors. Several antivirals are now being designed to block attachment of transcription factors to viral DNA.

Translation/antisense

Genomics has not only helped find targets for many antivirals, it has provided the basis for an entirely new type of drug, based on "antisense" molecules. These are segments of DNA or RNA that are designed as complementary molecule to critical sections of viral genomes, and the binding of these antisense segments to these target sections blocks the operation of those genomes. A phosphorothioate antisense drug named fomivirsen has been introduced, used to treat opportunistic eye infections in AIDS patients caused by cytomegalovirus, and other antisense antivirals are in development. An antisense structural type that has proven especially valuable in research is morpholino antisense.

Morpholino oligos have been used to experimentally suppress many viral types:

Translation/ribozymes

Yet another antiviral technique inspired by genomics is a set of drugs based on ribozymes, which are enzymes that will cut apart viral RNA or DNA at selected sites. In their natural course, ribozymes are used as part of the viral manufacturing sequence, but these synthetic ribozymes are designed to cut RNA and DNA at sites that will disable them.

A ribozyme antiviral to deal with hepatitis C has been suggested, [23] and ribozyme antivirals are being developed to deal with HIV. [24] An interesting variation of this idea is the use of genetically modified cells that can produce custom-tailored ribozymes. This is part of a broader effort to create genetically modified cells that can be injected into a host to attack pathogens by generating specialized proteins that block viral replication at various phases of the viral life cycle.

Protein processing and targeting

Interference with post translational modifications or with targeting of viral proteins in the cell is also possible. [25]

Protease inhibitors

Some viruses include an enzyme known as a protease that cuts viral protein chains apart so they can be assembled into their final configuration. HIV includes a protease, and so considerable research has been performed to find "protease inhibitors" to attack HIV at that phase of its life cycle. [26] Protease inhibitors became available in the 1990s and have proven effective, though they can have unusual side effects, for example causing fat to build up in unusual places. [27] Improved protease inhibitors are now in development.

Protease inhibitors have also been seen in nature. A protease inhibitor was isolated from the shiitake mushroom (Lentinus edodes). [28] The presence of this may explain the Shiitake mushrooms' noted antiviral activity in vitro. [29]

Long dsRNA helix targeting

Most viruses produce long dsRNA helices during transcription and replication. In contrast, uninfected mammalian cells generally produce dsRNA helices of fewer than 24 base pairs during transcription. DRACO (double-stranded RNA activated caspase oligomerizer) is a group of experimental antiviral drugs initially developed at the Massachusetts Institute of Technology. In cell culture, DRACO was reported to have broad-spectrum efficacy against many infectious viruses, including dengue flavivirus, Amapari and Tacaribe arenavirus, Guama bunyavirus, H1N1 influenza and rhinovirus, and was additionally found effective against influenza in vivo in weanling mice. It was reported to induce rapid apoptosis selectively in virus-infected mammalian cells, while leaving uninfected cells unharmed. [30] DRACO effects cell death via one of the last steps in the apoptosis pathway in which complexes containing intracellular apoptosis signalling molecules simultaneously bind multiple procaspases. The procaspases transactivate via cleavage, activate additional caspases in the cascade, and cleave a variety of cellular proteins, thereby killing the cell.[ citation needed ]

Assembly

Rifampicin acts at the assembly phase. [31]

Release phase

The final stage in the life cycle of a virus is the release of completed viruses from the host cell, and this step has also been targeted by antiviral drug developers. Two drugs named zanamivir (Relenza) and oseltamivir (Tamiflu) that have been recently introduced to treat influenza prevent the release of viral particles by blocking a molecule named neuraminidase that is found on the surface of flu viruses, and also seems to be constant across a wide range of flu strains.

Immune system stimulation

Rather than attacking viruses directly, a second category of tactics for fighting viruses involves encouraging the body's immune system to attack them. Some antivirals of this sort do not focus on a specific pathogen, instead stimulating the immune system to attack a range of pathogens.

One of the best-known of this class of drugs are interferons, which inhibit viral synthesis in infected cells. [32] One form of human interferon named "interferon alpha" is well-established as part of the standard treatment for hepatitis B and C, [33] and other interferons are also being investigated as treatments for various diseases.

A more specific approach is to synthesize antibodies, protein molecules that can bind to a pathogen and mark it for attack by other elements of the immune system. Once researchers identify a particular target on the pathogen, they can synthesize quantities of identical "monoclonal" antibodies to link up that target. A monoclonal drug is now being sold to help fight respiratory syncytial virus in babies, [34] and antibodies purified from infected individuals are also used as a treatment for hepatitis B. [35]

Antiviral drug resistance

Antiviral resistance can be defined by a decreased susceptibility to a drug caused by changes in viral genotypes. In cases of antiviral resistance, drugs have either diminished or no effectiveness against their target virus. [36] The issue inevitably remains a major obstacle to antiviral therapy as it has developed to almost all specific and effective antimicrobials, including antiviral agents. [37]

The Centers for Disease Control and Prevention (CDC) inclusively recommends anyone six months and older to get a yearly vaccination to protect them from influenza A viruses (H1N1) and (H3N2) and up to two influenza B viruses (depending on the vaccination). [36] Comprehensive protection starts by ensuring vaccinations are current and complete. However, vaccines are preventative and are not generally used once a patient has been infected with a virus. Additionally, the availability of these vaccines can be limited based on financial or locational reasons which can prevent the effectiveness of herd immunity, making effective antivirals a necessity. [36]

The three FDA-approved neuraminidase antiviral flu drugs available in the United States, recommended by the CDC, include: oseltamivir (Tamiflu), zanamivir (Relenza), and peramivir (Rapivab). [36] Influenza antiviral resistance often results from changes occurring in neuraminidase and hemagglutinin proteins on the viral surface. Currently, neuraminidase inhibitors (NAIs) are the most frequently prescribed antivirals because they are effective against both influenza A and B. However, antiviral resistance is known to develop if mutations to the neuraminidase proteins prevent NAI binding. [38] This was seen in the H257Y mutation, which was responsible for oseltamivir resistance to H1N1 strains in 2009. [36] The inability of NA inhibitors to bind to the virus allowed this strain of virus with the resistance mutation to spread due to natural selection. Furthermore, a study published in 2009 in Nature Biotechnology emphasized the urgent need for augmentation of oseltamivir stockpiles with additional antiviral drugs including zanamivir. This finding was based on a performance evaluation of these drugs supposing the 2009 H1N1 'Swine Flu' neuraminidase (NA) were to acquire the oseltamivir-resistance (His274Tyr) mutation, which is currently widespread in seasonal H1N1 strains. [39]

Origin of antiviral resistance

The genetic makeup of viruses is constantly changing, which can cause a virus to become resistant to currently available treatments. [40] Viruses can become resistant through spontaneous or intermittent mechanisms throughout the course of an antiviral treatment. [36] Immunocompromised patients, more often than immunocompetent patients, hospitalized with pneumonia are at the highest risk of developing oseltamivir resistance during treatment. [36] Subsequent to exposure to someone else with the flu, those who received oseltamivir for "post-exposure prophylaxis" are also at higher risk of resistance. [41]

The mechanisms for antiviral resistance development depend on the type of virus in question. RNA viruses such as hepatitis C and influenza A have high error rates during genome replication because RNA polymerases lack proofreading activity. [42] RNA viruses also have small genome sizes that are typically less than 30 kb, which allow them to sustain a high frequency of mutations. [43] DNA viruses, such as HPV and herpesvirus, hijack host cell replication machinery, which gives them proofreading capabilities during replication. DNA viruses are therefore less error prone, are generally less diverse, and are more slowly evolving than RNA viruses. [42] In both cases, the likelihood of mutations is exacerbated by the speed with which viruses reproduce, which provides more opportunities for mutations to occur in successive replications. Billions of viruses are produced every day during the course of an infection, with each replication giving another chance for mutations that encode for resistance to occur. [44]

Multiple strains of one virus can be present in the body at one time, and some of these strains may contain mutations that cause antiviral resistance. [37] This effect, called the quasispecies model, results in immense variation in any given sample of virus, and gives the opportunity for natural selection to favor viral strains with the highest fitness every time the virus is spread to a new host. [45] Recombination, the joining of two different viral variants, and reassortment, the swapping of viral gene segments among viruses in the same cell, also play a role in resistance, especially in influenza. [43]

Antiviral resistance has been reported in antivirals for herpes, HIV, hepatitis B and C, and influenza, but antiviral resistance is a possibility for all viruses. [37] Mechanisms of antiviral resistance vary between virus types.[ citation needed ]

Detection of antiviral resistance

National and international surveillance is performed by the CDC to determine effectiveness of the current FDA-approved antiviral flu drugs. [36] Public health officials use this information to make current recommendations about the use of flu antiviral medications. WHO further recommends in-depth epidemiological investigations to control potential transmission of the resistant virus and prevent future progression. [46] As novel treatments and detection techniques to antiviral resistance are enhanced so can the establishment of strategies to combat the inevitable emergence of antiviral resistance. [47]

Treatment options for antiviral resistant pathogens

If a virus is not fully wiped out during a regimen of antivirals, treatment creates a bottleneck in the viral population that selects for resistance, and there is a chance that a resistant strain may repopulate the host. [48] Viral treatment mechanisms must therefore account for the selection of resistant viruses.

The most commonly used method for treating resistant viruses is combination therapy, which uses multiple antivirals in one treatment regimen. This is thought to decrease the likelihood that one mutation could cause antiviral resistance, as the antivirals in the cocktail target different stages of the viral life cycle. [49] This is frequently used in retroviruses like HIV, but a number of studies have demonstrated its effectiveness against influenza A, as well. [50] Viruses can also be screened for resistance to drugs before treatment is started. This minimizes exposure to unnecessary antivirals and ensures that an effective medication is being used. This may improve patient outcomes and could help detect new resistance mutations during routine scanning for known mutants. [48] However, this has not been consistently implemented in treatment facilities at this time.

Vaccinations

While most antivirals treat viral infection, vaccines are a preemptive first line of defense against pathogens. Vaccination involves the introduction (i.e. via injection) of a small amount of typically inactivated or attenuated antigenic material to stimulate an individual's immune system. The immune system responds by developing white blood cells to specifically combat the introduced pathogen, resulting in adaptive immunity. [51] Vaccination in a population results in herd immunity and greatly improved population health, with significant reductions in viral infection and disease. [52]

Vaccination policy

Vaccination policy in the United States consists of public and private vaccination requirements. For instance, public schools require students to receive vaccinations (termed "vaccination schedule") for viruses and bacteria such as diphtheria, pertussis, and tetanus (DTaP), measles, mumps, rubella (MMR), varicella (chickenpox), hepatitis B, rotavirus, polio, and more. Private institutions might require annual influenza vaccination. The Centers for Disease Control and Prevention has estimated that routine immunization of newborns prevents about 42,000 deaths and 20 million cases of disease each year, saving about $13.6 billion. [53]

Vaccination controversy

Despite their successes, in the United States there exists plenty of stigma surrounding vaccines that cause people to be incompletely vaccinated. These "gaps" in vaccination result in unnecessary infection, death, and costs. [54] There are two major reasons for incomplete vaccination:

  1. Vaccines, like other medical treatments, have a risk of causing complications in some individuals (allergic reactions). Vaccines do not cause autism; this has been confirmed by national health agencies, such as the US Centers for Disease Control and Prevention, [55] the US Institute of Medicine, [56] and the UK National Health Service [57]
  2. Low rates of vaccine-preventable disease, as a result of herd immunity, also make vaccines seem unnecessary and leave many unvaccinated. [58] [59]

Although the American Academy of Pediatrics endorses universal immunization, [60] they note that physicians should respect parents' refusal to vaccinate their children after sufficient advising and provided the child does not face a significant risk of infection. Parents can also cite religious reasons to avoid public school vaccination mandates, but this reduces herd immunity and increases risk of viral infection. [52]

Limitations of vaccines

Vaccines boosts the body's immune system to better attack viruses in the "complete particle" stage, outside of the organism's cells. Traditional approaches to vaccine development include an attenuated (a live weakened) or inactivated (killed) version of the virus. Attenuated pathogens, in very rare cases, can revert to a pathogenic form. Inactivated vaccines have no possibility of introducing the disease they are given against; on the other hand, the immune response may not always occur or it may be short lived, requiring several doses. Recently, "subunit" vaccines have been devised containing only the antigenic parts of the pathogen. This makes the vaccine "more precise" but without guarantee that immunological memory will be formed in the correct manner. [61]

Vaccines are very effective on stable viruses but are of limited use in treating a patient who has already been infected. They are also difficult to successfully deploy against rapidly mutating viruses, such as influenza (the vaccine for which is updated every year) and HIV. Antiviral drugs are particularly useful in these cases.

Antiretroviral therapy as HIV prevention

Following the HPTN 052 study and PARTNER study, there is significant evidence to demonstrate that antiretroviral drugs inhibit transmission when the HIV virus in the person living with HIV has been undetectable for 6 months or longer. [62] [63]

Public policy

Use and distribution

Guidelines regarding viral diagnoses and treatments change frequently and limit quality care. [64] Even when physicians diagnose older patients with influenza, use of antiviral treatment can be low. [65] Provider knowledge of antiviral therapies can improve patient care, especially in geriatric medicine. Furthermore, in local health departments (LHDs) with access to antivirals, guidelines may be unclear, causing delays in treatment. [66] With time-sensitive therapies, delays could lead to lack of treatment. Overall, national guidelines, regarding infection control and management, standardize care and improve healthcare worker and patient safety. Guidelines, such as those provided by the Centers for Disease Control and Prevention (CDC) during the 2009 flu pandemic caused by the H1N1 virus, recommend, among other things, antiviral treatment regimens, clinical assessment algorithms for coordination of care, and antiviral chemoprophylaxis guidelines for exposed persons. [67] Roles of pharmacists and pharmacies have also expanded to meet the needs of public during public health emergencies. [68]

Stockpiling

Public Health Emergency Preparedness initiatives are managed by the CDC via the Office of Public Health Preparedness and Response. [69] Funds aim to support communities in preparing for public health emergencies, including pandemic influenza. Also managed by the CDC, the Strategic National Stockpile (SNS) consists of bulk quantities of medicines and supplies for use during such emergencies. [70] Antiviral stockpiles prepare for shortages of antiviral medications in cases of public health emergencies. During the H1N1 pandemic in 2009–2010, guidelines for SNS use by local health departments was unclear, revealing gaps in antiviral planning. [66] For example, local health departments that received antivirals from the SNS did not have transparent guidance on the use of the treatments. The gap made it difficult to create plans and policies for their use and future availabilities, causing delays in treatment.

See also

Related Research Articles

<span class="mw-page-title-main">Hepatitis</span> Inflammation of the liver

Hepatitis is inflammation of the liver tissue. Some people or animals with hepatitis have no symptoms, whereas others develop yellow discoloration of the skin and whites of the eyes (jaundice), poor appetite, vomiting, tiredness, abdominal pain, and diarrhea. Hepatitis is acute if it resolves within six months, and chronic if it lasts longer than six months. Acute hepatitis can resolve on its own, progress to chronic hepatitis, or (rarely) result in acute liver failure. Chronic hepatitis may progress to scarring of the liver (cirrhosis), liver failure, and liver cancer.

The management of HIV/AIDS normally includes the use of multiple antiretroviral drugs as a strategy to control HIV infection. There are several classes of antiretroviral agents that act on different stages of the HIV life-cycle. The use of multiple drugs that act on different viral targets is known as highly active antiretroviral therapy (HAART). HAART decreases the patient's total burden of HIV, maintains function of the immune system, and prevents opportunistic infections that often lead to death. HAART also prevents the transmission of HIV between serodiscordant same-sex and opposite-sex partners so long as the HIV-positive partner maintains an undetectable viral load.

Gene silencing is the regulation of gene expression in a cell to prevent the expression of a certain gene. Gene silencing can occur during either transcription or translation and is often used in research. In particular, methods used to silence genes are being increasingly used to produce therapeutics to combat cancer and other diseases, such as infectious diseases and neurodegenerative disorders.

<span class="mw-page-title-main">Respiratory syncytial virus</span> Species of a virus

Respiratory syncytial virus (RSV), also called human respiratory syncytial virus (hRSV) and human orthopneumovirus, is a common, contagious virus that causes infections of the respiratory tract. It is a negative-sense, single-stranded RNA virus. Its name is derived from the large cells known as syncytia that form when infected cells fuse.

<span class="mw-page-title-main">Rimantadine</span> Drug used to treat influenzavirus A infection

Rimantadine is an orally administered antiviral drug used to treat, and in rare cases prevent, influenzavirus A infection. When taken within one to two days of developing symptoms, rimantadine can shorten the duration and moderate the severity of influenza. Rimantadine can mitigate symptoms, including fever. Both rimantadine and the similar drug amantadine are derivates of adamantane. Rimantadine is found to be more effective than amantadine because when used the patient displays fewer symptoms. Rimantadine was approved by the Food and Drug Administration (FDA) in 1994.

<span class="mw-page-title-main">Influenza treatment</span> Therapy and pharmacy for the common infectious disease

Treatments for influenza include a range of medications and therapies that are used in response to disease influenza. Treatments may either directly target the influenza virus itself; or instead they may just offer relief to symptoms of the disease, while the body's own immune system works to recover from infection.

<span class="mw-page-title-main">Umifenovir</span> Chemical compound

Umifenovir, sold under the brand name Arbidol, is an antiviral medication for the treatment of influenza and COVID infections used in Russia and China. The drug is manufactured by Pharmstandard. It is not approved by the U.S. Food and Drug Administration (FDA) for the treatment or prevention of influenza.

Entry inhibitors, also known as fusion inhibitors, are a class of antiviral drugs that prevent a virus from entering a cell, for example, by blocking a receptor. Entry inhibitors are used to treat conditions such as HIV and hepatitis D.

<i>Human betaherpesvirus 5</i> Species of virus

Human betaherpesvirus 5, also called human cytomegalovirus (HCMV), is species of virus in the genus Cytomegalovirus, which in turn is a member of the viral family known as Herpesviridae or herpesviruses. It is also commonly called CMV. Within Herpesviridae, HCMV belongs to the Betaherpesvirinae subfamily, which also includes cytomegaloviruses from other mammals. CMV is a double-stranded DNA virus.

<span class="mw-page-title-main">Nitazoxanide</span> Broad-spectrum antiparasitic and antiviral medication

Nitazoxanide, sold under the brand name Alinia among others, is a broad-spectrum antiparasitic and broad-spectrum antiviral medication that is used in medicine for the treatment of various helminthic, protozoal, and viral infections. It is indicated for the treatment of infection by Cryptosporidium parvum and Giardia lamblia in immunocompetent individuals and has been repurposed for the treatment of influenza. Nitazoxanide has also been shown to have in vitro antiparasitic activity and clinical treatment efficacy for infections caused by other protozoa and helminths; evidence as of 2014 suggested that it possesses efficacy in treating a number of viral infections as well.

<span class="mw-page-title-main">Resistance mutation (virology)</span> Virus mutation

A resistance mutation is a mutation in a virus gene that allows the virus to become resistant to treatment with a particular antiviral drug. The term was first used in the management of HIV, the first virus in which genome sequencing was routinely used to look for drug resistance. At the time of infection, a virus will infect and begin to replicate within a preliminary cell. As subsequent cells are infected, random mutations will occur in the viral genome. When these mutations begin to accumulate, antiviral methods will kill the wild type strain, but will not be able to kill one or many mutated forms of the original virus. At this point a resistance mutation has occurred because the new strain of virus is now resistant to the antiviral treatment that would have killed the original virus. Resistance mutations are evident and widely studied in HIV due to its high rate of mutation and prevalence in the general population. Resistance mutation is now studied in bacteriology and parasitology.

<span class="mw-page-title-main">Introduction to viruses</span> Non-technical introduction to viruses

A virus is a tiny infectious agent that reproduces inside the cells of living hosts. When infected, the host cell is forced to rapidly produce thousands of identical copies of the original virus. Unlike most living things, viruses do not have cells that divide; new viruses assemble in the infected host cell. But unlike simpler infectious agents like prions, they contain genes, which allow them to mutate and evolve. Over 4,800 species of viruses have been described in detail out of the millions in the environment. Their origin is unclear: some may have evolved from plasmids—pieces of DNA that can move between cells—while others may have evolved from bacteria.

<span class="mw-page-title-main">Virus</span> Infectious agent that replicates in cells

A virus is a submicroscopic infectious agent that replicates only inside the living cells of an organism. Viruses infect all life forms, from animals and plants to microorganisms, including bacteria and archaea. Viruses are found in almost every ecosystem on Earth and are the most numerous type of biological entity. Since Dmitri Ivanovsky's 1892 article describing a non-bacterial pathogen infecting tobacco plants and the discovery of the tobacco mosaic virus by Martinus Beijerinck in 1898, more than 11,000 of the millions of virus species have been described in detail. The study of viruses is known as virology, a subspeciality of microbiology.

<span class="mw-page-title-main">Influenza</span> Infectious disease, often just "the flu"

Influenza, commonly known as "the flu", is an infectious disease caused by influenza viruses. Symptoms range from mild to severe and often include fever, runny nose, sore throat, muscle pain, headache, coughing, and fatigue. These symptoms begin from one to four days after exposure to the virus and last for about 2–8 days. Diarrhea and vomiting can occur, particularly in children. Influenza may progress to pneumonia, which can be caused by the virus or by a subsequent bacterial infection. Other complications of infection include acute respiratory distress syndrome, meningitis, encephalitis, and worsening of pre-existing health problems such as asthma and cardiovascular disease.

<span class="mw-page-title-main">Viral neuraminidase</span>

Viral neuraminidase is a type of neuraminidase found on the surface of influenza viruses that enables the virus to be released from the host cell. Neuraminidases are enzymes that cleave sialic acid groups from glycoproteins. Viral neuraminidase was discovered by Alfred Gottschalk at the Walter and Eliza Hall Institute in 1957. Neuraminidase inhibitors are antiviral agents that inhibit influenza viral neuraminidase activity and are of major importance in the control of influenza.

A neutralizing antibody (NAb) is an antibody that defends a cell from a pathogen or infectious particle by neutralizing any effect it has biologically. Neutralization renders the particle no longer infectious or pathogenic. Neutralizing antibodies are part of the humoral response of the adaptive immune system against viruses, intracellular bacteria and microbial toxin. By binding specifically to surface structures (antigen) on an infectious particle, neutralizing antibodies prevent the particle from interacting with its host cells it might infect and destroy.

<span class="mw-page-title-main">Antiviral protein</span>

Antiviral proteins are proteins that are induced by human or animal cells to interfere with viral replication. These proteins are isolated to inhibit the virus from replicating in a host's cells and stop it from spreading to other cells. The Pokeweed antiviral protein and the Zinc-Finger antiviral protein are two major antiviral proteins that have undergone several tests for viruses, including HIV and influenza.

<span class="mw-page-title-main">HIV/AIDS research</span> Field of immunology research

HIV/AIDS research includes all medical research that attempts to prevent, treat, or cure HIV/AIDS, as well as fundamental research about the nature of HIV as an infectious agent and AIDS as the disease caused by HIV.

<span class="mw-page-title-main">Discovery and development of NS5A inhibitors</span>

Nonstructural protein 5A (NS5A) inhibitors are direct acting antiviral agents (DAAs) that target viral proteins, and their development was a culmination of increased understanding of the viral life cycle combined with advances in drug discovery technology. However, their mechanism of action is complex and not fully understood. NS5A inhibitors were the focus of much attention when they emerged as a part of the first curative treatment for hepatitis C virus (HCV) infections in 2014. Favorable characteristics have been introduced through varied structural changes, and structural similarities between NS5A inhibitors that are clinically approved are readily apparent. Despite the recent introduction of numerous new antiviral drugs, resistance is still a concern and these inhibitors are therefore always used in combination with other drugs.

HSV epigenetics is the epigenetic modification of herpes simplex virus (HSV) genetic code.

References

  1. Antiviral Agents. 2012. PMID   31643973.
  2. Rossignol JF (2014). "Nitazoxanide: a first-in-class broad-spectrum antiviral agent". Antiviral Res. 110: 94–103. doi: 10.1016/j.antiviral.2014.07.014 . PMC   7113776 . PMID   25108173.
  3. Rick Daniels; Leslie H. Nicoll. "Pharmacology – Nursing Management". Contemporary Medical-Surgical Nursing. Cengage Learning, 2011. p. 397.
  4. Ko K, Tekoah Y, Rudd PM, Harvey DJ, Dwek RA, Spitsin S, Hanlon CA, Rupprecht C, Dietzschold B, Golovkin M, Koprowski H (2003). "Function and glycosylation of plant-derived antiviral monoclonal antibody". PNAS. 100 (13): 8013–18. Bibcode:2003PNAS..100.8013K. doi: 10.1073/pnas.0832472100 . PMC   164704 . PMID   12799460.
  5. Schnitzler, P; Schön, K; Reichling, J (2001). "Antiviral activity of Australian tea tree oil and eucalyptus oil against herpes simplex virus in cell culture". Die Pharmazie. 56 (4): 343–47. PMID   11338678.
  6. Kausar S, Said Khan F, Ishaq Mujeeb Ur Rehman M, Akram M, Riaz M, Rasool G, Hamid Khan A, Saleem I, Shamim S, Malik A (2021). "A review: Mechanism of action of antiviral drugs". International Journal of Immunopathology and Pharmacology. 35: 20587384211002621. doi:10.1177/20587384211002621. PMC   7975490 . PMID   33726557.
  7. Yin H, Jiang N, Shi W, Chi X, Liu S, Chen JL, Wang S (February 2021). "Development and Effects of Influenza Antiviral Drugs". Molecules (Basel, Switzerland). 26 (4): 810. doi: 10.3390/molecules26040810 . PMC   7913928 . PMID   33557246.
  8. Bobrowski T, Melo-Filho CC, Korn D, Alves VM, Popov KI, Auerbach S, Schmitt C, Moorman NJ, Muratov EN, Tropsha A (September 2020). "Learning from history: do not flatten the curve of antiviral research!". Drug Discovery Today. 25 (9): 1604–1613. doi:10.1016/j.drudis.2020.07.008. PMC   7361119 . PMID   32679173.
  9. Lindequist, Ulrike; Niedermeyer, Timo H. J.; Jülich, Wolf-Dieter (2005). "The Pharmacological Potential of Mushrooms". Evidence-Based Complementary and Alternative Medicine. 2 (3): 285–99. doi:10.1093/ecam/neh107. PMC   1193547 . PMID   16136207.
  10. Pradeep, Prabin; Manju, Vidya; Ahsan, Mohammad Feraz (2019), Agrawal, Dinesh Chandra; Dhanasekaran, Muralikrishnan (eds.), "Antiviral Potency of Mushroom Constituents", Medicinal Mushrooms: Recent Progress in Research and Development, Springer Singapore, pp. 275–97, doi:10.1007/978-981-13-6382-5_10, ISBN   978-981-13-6382-5, S2CID   181538245
  11. Bishop NE (1998). "Examination of potential inhibitors of hepatitis A virus uncoating". Intervirology. 41 (6): 261–71. doi:10.1159/000024948. PMID   10325536. S2CID   21222121.
  12. Almela MJ, González ME, Carrasco L (May 1991). "Inhibitors of poliovirus uncoating efficiently block the early membrane permeabilization induced by virus particles". J. Virol. 65 (5): 2572–77. doi:10.1128/JVI.65.5.2572-2577.1991. PMC   240614 . PMID   1850030.
  13. Beringer, Paul; Troy, David A.; Remington, Joseph P. (2006). Remington, the science and practice of pharmacy. Hagerstwon, MD: Lippincott Williams & Wilkins. p. 1419. ISBN   978-0-7817-4673-1.
  14. Daniel C. Pevear; Tina M. Tull; Martin E. Seipel (1999). "Activity of Pleconaril against Enteroviruses". Antimicrobial Agents and Chemotherapy. 43 (9): 2109–2115. doi:10.1128/AAC.43.9.2109. PMC   89431 . PMID   10471549.
  15. Lee, S.; Nguyen, M.; Currier, M. (2016). "A polyvalent inactivated rhinovirus vaccine is broadly immunogenic in rhesus macaques". Nature Communications.
  16. "Common Cold Causes: Rhinoviruses and More". Archived from the original on 8 January 2022. Retrieved 8 January 2022.
  17. Tang, Roderick; Moore, Martin (2017). "Development of polyvalent inactivated rhinovirus vaccine".
  18. Stein DA, Skilling DE, Iversen PL, Smith AW (2001). "Inhibition of Vesivirus infections in mammalian tissue culture with antisense morpholino oligomers". Antisense Nucleic Acid Drug Dev. 11 (5): 317–25. doi:10.1089/108729001753231696. PMID   11763348.
  19. Deas, T. S.; Binduga-Gajewska, I.; Tilgner, M.; Ren, P.; Stein, D. A.; Moulton, H. M.; Iversen, P. L.; Kauffman, E. B.; Kramer, L. D.; Shi, P. -Y. (2005). "Inhibition of Flavivirus Infections by Antisense Oligomers Specifically Suppressing Viral Translation and RNA Replication". Journal of Virology. 79 (8): 4599–4609. doi:10.1128/JVI.79.8.4599-4609.2005. PMC   1069577 . PMID   15795246.
  20. Kinney, R. M.; Huang, C. Y.-H.; Rose, B. C.; Kroeker, A. D.; Dreher, T. W.; Iversen, P. L.; Stein, D. A. (2005). "Inhibition of Dengue Virus Serotypes 1 to 4 in Vero Cell Cultures with Morpholino Oligomers". J. Virol. 79 (8): 5116–28. doi:10.1128/JVI.79.8.5116-5128.2005. PMC   1069583 . PMID   15795296.
  21. McCaffrey AP, Meuse L, Karimi M, Contag CH, Kay MA (2003). "A potent and specific morpholino antisense inhibitor of hepatitis C translation in mice". Hepatology. 38 (2): 503–08. doi: 10.1053/jhep.2003.50330 . PMID   12883495. S2CID   1612244.
  22. Neuman, B. W.; Stein, D. A.; Kroeker, A. D.; Paulino, A. D.; Moulton, H. M.; Iversen, P. L.; Buchmeier, M. J. (June 2004). "Antisense Morpholino-Oligomers Directed against the 5' End of the Genome Inhibit Coronavirus Proliferation and Growth†". J. Virol. 78 (11): 5891–99. doi:10.1128/JVI.78.11.5891-5899.2004. PMC   415795 . PMID   15140987.
  23. Ryu KJ, Lee SW (2003). "Identification of the most accessible sites to ribozymes on the hepatitis C virus internal ribosome entry site". J. Biochem. Mol. Biol. 36 (6): 538–44. doi: 10.5483/BMBRep.2003.36.6.538 . PMID   14659071.
  24. Bai J, Rossi J, Akkina R (March 2001). "Multivalent anti-CCR ribozymes for stem cell-based HIV type 1 gene therapy". AIDS Res. Hum. Retroviruses. 17 (5): 385–99. doi:10.1089/088922201750102427. PMID   11282007.
  25. Alarcón B, González ME, Carrasco L (1988). "Megalomycin C, a macrolide antibiotic that blocks protein glycosylation and shows antiviral activity". FEBS Lett. 231 (1): 207–11. doi: 10.1016/0014-5793(88)80732-4 . PMID   2834223. S2CID   43114821.
  26. Anderson J, Schiffer C, Lee SK, Swanstrom R (2009). "Viral Protease Inhibitors". Antiviral Strategies. Handbook of Experimental Pharmacology. Vol. 189. pp. 85–110. doi:10.1007/978-3-540-79086-0_4. ISBN   978-3-540-79085-3. PMC   7120715 . PMID   19048198.
  27. Flint, O. P.; Noor, M. A.; Hruz, P. W.; Hylemon, P. B.; Yarasheski, K.; Kotler, D. P.; Parker, R. A.; Bellamine, A. (2009). "The Role of Protease Inhibitors in the Pathogenesis of HIV-Associated Lipodystrophy: Cellular Mechanisms and Clinical Implications". Toxicol Pathol. 37 (1): 65–77. doi:10.1177/0192623308327119. PMC   3170409 . PMID   19171928.
  28. Odani S, Tominaga K, Kondou S (1999). "The inhibitory properties and primary structure of a novel serine proteinase inhibitor from the fruiting body of the basidiomycete, Lentinus edodes". European Journal of Biochemistry. 262 (3): 915–23. doi: 10.1046/j.1432-1327.1999.00463.x . PMID   10411656.
  29. Suzuki H, Okubo A, Yamazaki S, Suzuki K, Mitsuya H, Toda S (1989). "Inhibition of the infectivity and cytopathic effect of human immunodeficiency virus by water-soluble lignin in an extract of the culture medium of Lentinus edodes mycelia (LEM)". Biochemical and Biophysical Research Communications. 160 (1): 367–73. doi:10.1016/0006-291X(89)91665-3. PMID   2469420.
  30. Rider TH, Zook CE, Boettcher TL, Wick ST, Pancoast JS, Zusman BD (2011). "Broad-spectrum antiviral therapeutics". PLOS ONE. 6 (7): e22572. Bibcode:2011PLoSO...622572R. doi: 10.1371/journal.pone.0022572 . PMC   3144912 . PMID   21818340.
  31. Sodeik B; Griffiths G; Ericsson M; Moss B; Doms RW (1994). "Assembly of vaccinia virus: effects of rifampin on the intracellular distribution of viral protein p65". J. Virol. 68 (2): 1103–14. doi:10.1128/JVI.68.2.1103-1114.1994. PMC   236549 . PMID   8289340.
  32. Samuel CE (October 2001). "Antiviral Actions of Interferons". Clin. Microbiol. Rev. 14 (4): 778–809. doi:10.1128/CMR.14.4.778-809.2001. PMC   89003 . PMID   11585785.
  33. Burra P (2009). "Hepatitis C". Semin. Liver Dis. 29 (1): 53–65. doi:10.1055/s-0029-1192055. PMID   19235659. S2CID   260319327.
  34. Nokes JD, Cane PA (December 2008). "New strategies for control of respiratory syncytial virus infection". Curr. Opin. Infect. Dis. 21 (6): 639–43. doi:10.1097/QCO.0b013e3283184245. PMID   18978532. S2CID   3065481.
  35. Akay S, Karasu Z (November 2008). "Hepatitis B immune globulin and HBV-related liver transplantation". Expert Opin Biol Ther (Submitted manuscript). 8 (11): 1815–22. doi:10.1517/14712598.8.11.1815. PMID   18847315. S2CID   71595650. Archived from the original on 3 October 2021. Retrieved 31 October 2018.
  36. 1 2 3 4 5 6 7 8 "Influenza Antiviral Drug Resistance| Seasonal Influenza (Flu)". CDC. 25 October 2018.
  37. 1 2 3 Pillay, D; Zambon, M (1998). "Antiviral Drug Resistance". BMJ. 317 (7159): 660–62. doi:10.1136/bmj.317.7159.660. PMC   1113839 . PMID   9728000.
  38. Moss, Ronald; Davey, Richard; Steigbigel, Roy; Fang, Fang (June 2010). "Targeting pandemic influenza: a primer on influenza antivirals and drug resistance". Journal of Antimicrobial Chemotherapy. 65 (6): 1086–93. doi: 10.1093/jac/dkq100 . PMID   20375034 . Retrieved 30 October 2018.
  39. Soundararajan, V; Tharakaraman, K; Raman, R; Raguram, S; Shriver, Z; Sasisekharan, V; Sasisekharan, R (June 2009). "Extrapolating from sequence--the 2009 H1N1 'swine' influenza virus". Nature Biotechnology. 27 (6): 510–13. doi:10.1038/nbt0609-510. PMID   19513050. S2CID   22710439.
  40. Nijhuis, M; van Maarseveen, NM; Boucher, CA (2009). "Antiviral Resistance and Impact on Viral Replication Capacity: Evolution of Viruses Under Antiviral Pressure Occurs in Three Phases". Antiviral Strategies. Handbook of Experimental Pharmacology. Vol. 189. pp. 299–320. doi:10.1007/978-3-540-79086-0_11. ISBN   978-3-540-79085-3. PMID   19048205.
  41. "WHO | Antiviral use and the risk of drug resistance". www.who.int. Archived from the original on 1 September 2014.
  42. 1 2 Lodish, H; Berk, A; Zipursky, S (2000). Molecular Cell Biology: Viruses – Structure, Function, and Uses. New York, New York: W. H. Freeman and Company. Retrieved 1 December 2018.
  43. 1 2 Racaniello, Vincent. "The error-prone ways of RNA synthesis". Virology Blog. Retrieved 1 December 2018.
  44. Thebaud, G; Chadeouf, J; Morelli, M; McCauley, J; Haydon, D (2010). "The relationship between mutation frequency and replication strategy in positive sense single-stranded RNA viruses". Proc. Biol. Sci. 277 (1682): 809–17. doi:10.1098/rspb.2009.1247. PMC   2842737 . PMID   19906671.
  45. "Viruses are models for embracing diversity". Nature Microbiology. 3 (4): 389. 2018. doi: 10.1038/s41564-018-0145-3 . PMID   29588540.
  46. Hayden, FG; de Jong, MD (1 January 2011). "Emerging influenza antiviral resistance threats". The Journal of Infectious Diseases. 203 (1): 6–10. doi:10.1093/infdis/jiq012. PMC   3086431 . PMID   21148489.
  47. Kimberlin, DW; Whitley, RJ (March 1996). "Antiviral resistance: mechanisms, clinical significance, and future implications". The Journal of Antimicrobial Chemotherapy. 37 (3): 403–21. doi: 10.1093/jac/37.3.403 . PMID   9182098.
  48. 1 2 Irwin, K; Renzette, N; Kowalik, T; Jensen, J (2016). "Antiviral drug resistance as an adaptive process". Virus Evolution. 2 (1): vew014. doi:10.1093/ve/vew014. PMC   5499642 . PMID   28694997.
  49. Moscona, A (2009). "Global transmission of oseltamivir-resistant influenza". New England Journal of Medicine. 360 (10): 953–56. doi: 10.1056/NEJMp0900648 . PMID   19258250. S2CID   205104988.
  50. Strasfeld, L; Chou, S (2010). "Antiviral Drug Resistance: Mechanisms and Clinical Implications". Infectious Disease Clinics of North America. 24 (2): 413–37. doi:10.1016/j.idc.2010.01.001. PMC   2871161 . PMID   20466277.
  51. "Understanding How Vaccines Work". Centers for Disease Control and Prevention. 17 August 2021. Retrieved 11 October 2021.
  52. 1 2 Heymann, D. L.; Aylward, R. B. (2006). Mass vaccination: When and why. Vol. 304. pp. 1–16. doi:10.1007/3-540-36583-4_1. ISBN   978-3-540-29382-8. PMID   16989261. S2CID   25259803.{{cite book}}: |journal= ignored (help)
  53. Seither, R.; Masalovich, S.; Knighton, C. L.; Mellerson, J.; Singleton, J. A.; Greby, S. M.; Centers for Disease Control and Prevention (CDC) (2014). "Vaccination Coverage Among Children in Kindergarten—United States, 2013–14 School Year". MMWR. 63 (41): 913–920. PMC   4584748 . PMID   25321068.
  54. Omer, SB; Salmon, DA; Orenstein, WA; deHart, MP; Halsey, N (May 2009). "Vaccine Refusal, Mandatory Immunization, and the Risks of Vaccine-Preventable Diseases". New England Journal of Medicine. 360 (19): 1981–88. doi: 10.1056/NEJMsa0806477 . PMID   19420367. S2CID   5353949.
  55. "Vaccines Do Not Cause Autism". Centers for Disease Control and Prevention. 23 November 2015. Retrieved 20 October 2016.
  56. Immunization Safety Review Committee (2004).Immunization Safety Review: Vaccines and Autism. The National Academies Press. ISBN   0-309-09237-X.
  57. "MMR vaccine". National Health Service . Retrieved 20 October 2016.
  58. Hendriksz T, Malouf PH, Sarmiento S, Foy JE. "Overcoming patient barriers to immunizations". AOA Health Watch. 2013: 9–14.
  59. "Barriers and Strategies to Improving Influenza Vaccination among Health Care Personnel". Centers for Disease Control and Prevention. 7 September 2016. Retrieved 17 September 2016.
  60. Diekema DS (2005). "Responding to parental refusals of immunization of children". Pediatrics. 115 (5): 1428–31. doi: 10.1542/peds.2005-0316 . PMID   15867060.
  61. "Types of vaccine and adverse reactions" (PDF).
  62. "HPTN 052". HPTN. Retrieved 29 September 2017.
  63. Rodger AJ, Cambiano V, Bruun T, Vernazza P, Collins S, van Lunzen J, Corbelli GM, Estrada V, Geretti AM, Beloukas A, Asboe D, Viciana P, Gutiérrez F, Clotet B, Pradier C, Gerstoft J, Weber R, Westling K, Wandeler G, Prins JM, Rieger A, Stoeckle M, Kümmerle T, Bini T, Ammassari A, Gilson R, Krznaric I, Ristola M, Zangerle R, Handberg P, Antela A, Allan S, Phillips AN, Lundgren J (12 July 2016). "Sexual Activity Without Condoms and Risk of HIV Transmission in Serodifferent Couples When the HIV-Positive Partner Is Using Suppressive Antiretroviral Therapy". JAMA. 316 (2): 171–181. doi: 10.1001/jama.2016.5148 . PMID   27404185.
  64. Kunin, Marina; Engelhard, Dan; Thomas, Shane; Ashworth, Mark; Piterman, Leon (15 October 2015). "Challenges of the Pandemic Response in Primary Care during Pre-Vaccination Period: A Qualitative Study". Israel Journal of Health Policy Research. 4 (1): 32. doi: 10.1186/s13584-015-0028-5 . PMC   4606524 . PMID   26473026.
  65. Lindegren, Mary Louise; Griffin, Marie R.; Williams, John V.; Edwards, Kathryn M.; Zhu, Yuwei; Mitchel, Ed; Fry, Alicia M.; Schaffner, William; Talbot, H. Keipp; Pyrc, Krzysztof (25 March 2015). "Antiviral Treatment among Older Adults Hospitalized with Influenza, 2006–2012". PLOS ONE. 10 (3): e0121952. Bibcode:2015PLoSO..1021952L. doi: 10.1371/journal.pone.0121952 . PMC   4373943 . PMID   25807314.
  66. 1 2 NACCHO (December 2010). Public Health Use and Distribution of Antivirals: NACCHO Think Tank Meeting Report (PDF) (Report). Archived from the original (PDF) on 22 October 2016. Retrieved 21 October 2016.
  67. Centers for Disease Control and Prevention. "H1N1 Flu".
  68. Hodge, J G; Orenstein, D G. "Antiviral Distribution and Dispensing A Review of Legal and Policy Issues". Association of State and Territorial Health Officials (ASTHO).
  69. "Funding and Guidance for State and Local Health Departments". Centers for Disease Control and Prevention. Retrieved 21 October 2016.
  70. "Strategic National Stockpile (SNS)". Centers for Disease Control and Prevention. Retrieved 21 October 2016.