Announcement

Collapse
No announcement yet.

Cell-mediated Protection in Influenza Infection

Collapse
X
 
  • Filter
  • Time
  • Show
Clear All
new posts

  • Cell-mediated Protection in Influenza Infection



    Cell-mediated Protection in Influenza Infection

    Paul G. Thomas,* Rachael Keating,* Diane J. Hulse-Post,* and Peter C. Doherty*<sup></sup>
    *St. Jude Children's Research Hospital, Memphis, Tennessee, USA
    Suggested citation for this article
    <hr>
    Current vaccine strategies against influenza focus on generating robust antibody responses. Because of the high degree of antigenic drift among circulating influenza strains over the course of a year, vaccine strains must be reformulated specifically for each influenza season. The time delay from isolating the pandemic strain to large-scale vaccine production would be detrimental in a pandemic situation. A vaccine approach based on cell-mediated immunity that avoids some of these drawbacks is discussed here. Specifically, cell-mediated responses typically focus on peptides from internal influenza proteins, which are far less susceptible to antigenic variation. We review the literature on the role of CD4+ and CD8+ T cell?mediated immunity in influenza infection and the available data on the role of these responses in protection from highly pathogenic influenza infection. We discuss the advantages of developing a vaccine based on cell-mediated immune responses toward highly pathogenic influenza virus and potential problems arising from immune pressure.
    Vaccine approaches against respiratory virus infections such as influenza have relied on inducing antibodies that protect against viral infection by neutralizing virions or blocking the virus's entry into cells. These humoral immune responses target external viral coat proteins that are conserved for a given strain. Antibody-mediated protection is therefore effective against homologous viral strains but inadequate against heterologous strains with serologically distinct coat proteins. This distinction is of consequence since many viruses rapidly mutate their coat proteins; an effective humoral response?based vaccine against a form of the virus may be ineffective against next season's variant. In contrast, T cells, which mediate cellular immune responses, can target internal proteins common to heterologous viral strains. This property gives vaccines that induce protective cellular immune responses the potential to protect against heterologous viral strains.
    Antigen-specific ligation of T-cell receptors induces effector mechanisms that either directly or indirectly promote lysis of infected cells. Functionally distinct T-cell subsets are broadly identified according to their differential expression of CD4 and CD8 coreceptors. The CD4+ T helper cells are primarily responsible for helping other immune cells through direct cell-cell interactions or by secreting cytokines after recognizing viral peptides bound to major histocompatibility complex (MHC) class II molecules. The cytotoxic T lymphocytes (CTLs) typically express CD8 and induce apoptosis of cells on which they recognize foreign antigens presented by MHC class I molecules, providing a defense against intracellular pathogens such as viruses. This association of phenotype and function is not absolute, since CD4+ cells may exhibit lytic activity, while CD8+ cells secrete antiviral cytokines, notably interferon-γ (IFN-γ) and tumor necrosis factor. Greater understanding of how each subset contributes to protective immunity and how T-cell memory is maintained and recalled in a secondary infection would contribute to development of effective vaccines that use these basic features of the immune response.
    Immune Models of Influenza

    Influenza is a contagious, acute respiratory disease caused by infection of the host respiratory tract mucosa by an influenza virus (1). The influenza A viruses infect host epithelial cells by attaching to a cellular receptor (sialic acid) by the viral surface protein hemagglutinin (HA). The virus is subsequently released because of the action of another surface glycoprotein, the enzyme neuraminidase (NA), several hours after infection.
    Mouse models of influenza A virus pneumonia provide a well-developed experimental system to analyze T cell?mediated immunity. In particular, the T-cell immune response to influenza infection has been well characterized in C57BL/6 (B6,H2<sup>b</sup>) mice. While influenza infection of mice does not precisely replicate the natural infection in human, avian, or other vertebrate species, the availability of reagents and genetically modified mouse models has enabled extensive analysis of the cellular immune response. Emerging evidence indicates that findings from mouse studies are pertinent to immunopathology in human disease. In the BL/6 model, virus is cleared 10 days after infection, with no indication of persistent antigen or viral RNA (2). Recovery or prevention of influenza relies on targeting both innate and adaptive responses to the respiratory tract mucosa.
    CD8+ T-cell Response to Influenza

    Much of the current knowledge on murine CD8+ T-cell responses to influenza has come from analyzing the response to challenge with the HKx31 (H3N2) and PR/8 (H1N1) influenza viruses. A role for CD8+ T cells in protective immunity has been discerned from studies citing delayed influenza virus clearance in CD8+ T cell?deficient mice (3,4). Furthermore, CD8+ T cells can promote recovery from otherwise lethal secondary viral infections in mice that lack mature B cells or antibodies (5,6), and cloned influenza-specific CTLs can passively transfer protection (7). Despite a seemingly protective role for CD8+ T cells, vaccination with dominant influenza determinants in either a vector or in a recombinant form is only mildly protective (8?10). Moreover, in a T cell?receptor transgenic mouse model, devoid of antibodies, influenza-specific CTL can either contribute to survival or exacerbate lethal influenza pneumonia (11). This study highlights the need to understand the many facets of the immune response to influenza.
    The influenza A virus?specific CD8+ T-cell response has been characterized by using intracellular cytokine staining and MHC class I tetramer labeling. These techniques have enabled each phase of the response to be tracked. After intranasal infection, priming, activation, and expansion of naive influenza-specific CD8+ T cells occur in the draining mediastinal lymph node 3?4 days after infection (12,13). The antiviral capacity of these virus-specific CD8+ cells is strongly dependent on their ability to migrate and localize to the lungs and infected airway epithelium (14), where they appear 5?7 days after infection (15). Because viral replication is confined to cells in the respiratory epithelium (16,17), CD8+ T cells exert their effector functions at this site, producing antiviral cytokines and lysing target cells presenting viral determinants for which they bear a specific T-cell receptor. Lysis of infected epithelial cells is mediated by exocytosis granules containing perforin and granzymes (18,19). The release of perforin and granzymes from influenza-specific CTLs is tightly regulated, occurring shortly after activation at or near the contact point between CTLs and the infected target cell (18).
    Influenza-specific CD8+ T cells recognize multiple viral epitopes on target cells and antigen-presenting cells. The HKx31 and PR8 strains share 6 internal genes derived from PR8 that are processed to generate peptides recognized by influenza-specific CD8+ T cells. The primary response to either strain is dominated by CD8+ T cells' recognition of 2 determinants, the nucleoprotein (NP<sub>366-374</sub>, H2D<sup>b</sup>) (20) and the acid polymerase (PA<sub>224-233</sub>, H2D<sup>b</sup>) (21). A similarly low proportion of CD8+ T cells recognizes 4 other determinants: the basic polymerase subunit 1 (PB1<sub>703-711</sub>, H2K<sup>b</sup>), nonstructural protein 2 (NS2<sub>114-121</sub>, H2K<sup>d</sup>), matrix protein 1 (M1<sub>128-135</sub>, H2K<sup>b</sup>), and a protein derived from an alternative open reading frame within the PB1 gene (PB1-F2<sub>62-70</sub>, H2D<sup>b</sup>) (22). The subsequent memory populations appears to be stable; D<sup>b</sup>NP<sub>366-374</sub> and D<sup>b</sup>PA<sub>224-233</sub> CD8+ memory cells are still detectable >570 days after initial infection (K. Kedzierska and J. Stambas, unpub. data).
    Secondary influenza-specific CTL responses arise ≈2 days faster than the primary response, with a greatly increased level of activity. Depletion of CD8+ T cells reduces the capacity of primed mice to respond to influenza infection, which indicates a role for CD8+ T cells in the protective secondary response. Prime and challenge experiments can be conducted with HKx31 and PR/8 as all of the recognized epitopes are derived from internal proteins. Furthermore, cross-reactive neutralizing antibodies are avoided because HKx31 and PR/8 express different surface HA and NA or proteins. Despite a similar magnitude to D<sup>b</sup>PA<sub>224-233</sub> in the primary response, D<sup>b</sup>NP<sub>366-374</sub>-specific CD8+ T cells dominate the secondary response to HKx31→PR/8 challenge, accounting for up to 80% of the influenza-specific CD8+ T cells. This dominance is maintained in the memory population; the numbers of NP-specific CD8+ T cells exceed all other quantified influenza-specific CD8+ T-cell populations (23). Despite the NP dominance, CD8+ T cells specific for the other 5 determinants can still be isolated after secondary challenge, albeit at low frequency.
    Conservation of these 6 internal genes and persistence of the corresponding antigen-specific CD8+ T cells makes these genes an attractive target for vaccine therapies. However, although cell-mediated immunity can promote viral clearance, it does not provide sterile resistance because, unlike humoral immunity, it cannot prevent infection of the host cells. In humans, the level of influenza-specific CTLs correlates with the rate of viral clearance but not with susceptibility to infection or subsequent illness (24). Despite this limitation, vaccines that promote cell-mediated immunity may be a favorable option to fight potentially lethal, highly pathogenic influenza strains.
    CD4+ T cell?specific Responses to Viruses

    In contrast to the body of literature that has characterized the role of CD8+ T cells specifically in models of influenza infection, relatively little is known about the role of CD4+ T cells as direct mediators of effector function. That CD4+ T-cell help is central to adaptive immunity is well established, but few antigen-specific systems have been developed to dissect the role of CD4+ T cells in a viral infection. While knowledge of CD8+ T-cell antigen-specific responses has increased substantially in the past several years as a result of tetramer technology, these reagents have been more difficult to develop for the CD4+ subset. Further, identification of CD4+ T cell?specific epitopes has been less successful for a variety of pathogens. For instance, in influenza, the CD8+ restricted epitopes have all been largely identified for some time, particularly in the BL/6 model system; in contrast, only very recently have confirmed CD4 epitopes been found, and they are much more poorly characterized (25).
    Still, a substantial amount of work has been done with various knockout, depletion, and cell-transfer models to investigate the role of CD4+ T cells in primary, secondary, and memory responses to influenza infection in the mouse model (26,27). Controversy still exists in the field, and an antigen-specific system would help address it, but certain findings appear to be consistent across different experimental systems (28).
    In the primary response, CD4+ T cells are not required for expansion or development of functional CD8+ CTL (27,29), which may in part result from the ability of influenza virus to directly activate dendritic cells, aiding in the development of CD8+ responses that substitute for functional CD4+ T cells (30). Similarly, in the case of a murine γ-herpesvirus, the lack of CD4+ T cells can be compensated for by the addition of anti-CD40 stimulation (31). In mice in which both the CD4+ T-cell and B-cell compartments were defective, the primary CD8+ T-cell response to influenza appeared to be stunted in terms of recruitment and expansion (vs. mice in which B cells alone were knocked out); the remaining CD8+ T cells had a robust level of functionality as assayed by IFN-γ intracellular cytokine production (27). The defect in the CD8+ T-cell primary response was less obvious in mice with intact B cells, though viral clearance was delayed. Still, not until the secondary and memory responses are examined can the dramatic effect of CD4+ T-cell deletion be observed.
    In multiple systems, a defect of CD8+ T-cell secondary and memory responses have been observed when the primary response lacks CD4+ T cells (26,32,33). In influenza, a dramatic drop was observed in the size and magnitude of the recall response to secondary infection. The rate of viral clearance was also slowed considerably, beyond the degree seen in the primary response. Similarly, in the Listeria monocytogenes model system, the primary response was largely intact, while the long-term memory response was defective (34). In mice that lacked CD4+ T cells during the primary response, the memory pool of CD8+ T cells was initially similar in size and functionality to that seen in wild-type mice but began to decline after longer intervals, leading eventually to the recrudescence of the infection. Secondary challenge also demonstrated a reduced antigen-specific CD8+ T-cell compartment.
    In the influenza model, although the draining lymph node and spleen CD8+ responses were defective in secondary infection of CD4+ T cell?deficient mice, the CD8+ T-cell responses in bronchoalveolar lavage were equivalent to those seen in wild-type mice (29). This finding implies that the high levels of activation and inflammation, in large part mediated by innate immune effectors at the site of infection, were capable of providing the right maturation milieu to expand the response to wild-type levels; this finding suggests CD4+ T cell?specific help is not required at the site of the pathologic changes, at least when the infection induces a high level of other immune stimulation, though it is essential in the lymphoid organs in the generation and maintenance of memory.
    A role for CD4+ T cells as effectors has been found in a number of other systems, including the mouse γ-herpesvirus model (35) and in HIV-infected humans (36,37). In these studies, CD4+ T cells contribute to infection control by supplementing their helper role with cytotoxicity. In the case of the γ-herpesvirus, the effector CD4+ population was important only in immunoglobulin ?/? μMT mice, while the HIV studies were conducted in infected (and presumably immune-irregular) patients. However, effector CD4+ T cells have been found in multiple stages of the disease and in long-term patients whose disease is not progressing because viral replication is controlled. Finally, a recent report demonstrated a similar cytotoxic CD4+ T-cell effector population in protozoan-infected cattle (38).
    Relatively few established mouse models are available for studying the CD4+ response to influenza virus. On the IA<sup>d</sup> BALB/c background, an HA epitope was discovered, and a transgenic mouse was developed to analyze specific responses (the HNT model) (39). This model has been extremely useful for studying several aspects of CD4+ biology in influenza infection, particularly in regards to aging and the development of primary responses leading to acute memory (39). Several investigators have introduced external epitopes in influenza to follow CD4+ T-cell responses in defined systems. These include the hen egg lysozyme p46?63 sequence (40) and the ovalbumin 323?339 (OT-II) epitope inserted into the NA stalk of WSN influenza virus (41). We have inserted the OT-II epitope into the HA of the PR8 H1N1 virus and the X31 H3N2 virus. In contrast to the robust responses achieved with CD8+ T-cell epitopes and transgenics, the CD4+ T-cell responses seem relatively weak (unpub. data). Other naturally occurring epitopes have similarly low frequencies after infection (25). The antigen-specific CD4+ response may not develop the dramatic immunodominance hierarchies that are well-known for CD8+ T cells and may be directed at many epitopes, more than are seen in the more-delimited CD8+ T-cell response. Much work needs to be done before this conclusion is certain, and examples of respiratory infections in mice produce robust and dominant responses toward individual class II epitopes (42).
    Cell-mediated Protection against Highly Pathogenic Influenza

    Highly pathogenic H5N1 influenza emerged in 1997, followed by several waves of infection from 2002 until now (43). The viruses have been remarkably virulent in multiple animal models, including mice, but little work has been done to characterize the protective immune responses toward H5N1 viruses. A series of reports has shown strong protection toward other highly pathogenic viruses mediated by cellular responses, in the absence of neutralizing antibody. Antibody-deficient mice infected with a mild, passaged strain of an H3N2 virus were more likely to survive than naive controls when challenged with a highly pathogenic H3N8 duck virus compared to naive controls (44). A double-priming protocol provided increased protection from a lethal H7N7 challenge, which correlated with an increased pool of cross-reactive antigen-specific CD8+ T cells (45). In both these cases, the initial phase of infection and viral growth seemed similar to that occurring in immunologically naive mice, but a rapid decrease in viral titers occurs after a few days of infection.
    Since the emergence of the H5N1 viruses, concern has arisen that the biological activity of these viruses, including their diverse tissue tropism in a number of animal models, may influence the ability of immune responses to control infection. Furthermore, some pathology associated with these viruses has been attributed to extremely high levels of inflammatory cytokines produced in response to infection, which suggests a negative role for immune responses. However, the few studies that have been performed have shown promising results for the potential of cell-mediated responses to contribute to the control of infections. A prime-challenge protocol using an H9N2 isolate with 98% homology to the internal genes of the A/HongKong/156/97 H5N1 protected against the otherwise lethal challenge (46) with a virus with a highly cleavable H5, a characteristic of all the pathogenic H5 viruses. The priming protocol generated significant CTL activity directed at the NP and PB2 proteins.
    <table align="right" border="0" cellpadding="5" cellspacing="0" width="171"> <tbody><tr> <td width="11">
    </td> <td align="center" bgcolor="#fbed9d" width="150">Figure</td> <td align="center" width="10">
    </td> </tr> <tr> <td width="11">
    </td> <td align="center" bgcolor="#ffffcc" width="150"></td> <td align="center" width="10">
    </td> </tr> <tr> <td width="11">
    </td> <td bgcolor="#ffffcc" width="150"> Click to view enlarged image

    Figure. Apparent cell-mediated protection against highly pathogenic H5N1 influenza virus...
    </td> <td width="10">
    </td> </tr> </tbody></table> Our own work has indicated a similar ability of cell-mediated immunity to protect against virulent H5N1 challenge. In a preliminary experiment, we primed mice with the H1N1 PR8 strain and the H3N2 X31 strain followed by a challenge with A/Vietnam/1203/2004, one of the most lethal H5N1 viruses, which causes severe pathologic changes, even in ducks. While 9 of 10 naive mice died, 9 of 10 primed mice survived past day 10 of infectious challenge and recovered substantial weight (Figure). The fact that both groups lost weight indicated protection was occurring by delayed cell-mediated responses, rather than by the "immediate" cross-protective antibody response.
    Cell-mediated Vaccine for Highly Pathogenic Influenza?

    Despite the systems currently in place for manufacturing and distributing an influenza vaccine, pandemic influenza will require a substantially different approach. The standard influenza vaccine given during the infectious season is made from a reassortant seed strain containing the HA and NA of the circulating virus with the internal genes of a vaccine strain, usually PR8. The seed strain is grown in eggs and is formaldehyde inactivated. This strategy does not prime strong CD8 CTL responses, but it is effective in providing antibody-mediated protection to closely homologous strains (47).
    One drawback to this approach is the length of time required to develop a seed strain, amplify it, and manufacture it into distributable vaccine. In the case of a potential influenza pandemic, the delivery of vaccine on this schedule would not prevent the spread of the epidemic in many countries. Furthermore, antigenic drift can occur between the original selection of the seed strain and circulating viruses before the vaccine is ready for distribution (48). This problem was faced recently in a nonpandemic situation in 2003 and 2004 when the circulating Fujian strain of H3N2 influenza had drifted from the vaccine strain (49). While the Fujian strain was predicted to be circulating at the time of vaccine delivery, a recombined seed strain could not be isolated in time for vaccine production. Although the ensuing influenza season was not as severe as initially feared, the situation highlighted a problem with the current vaccine strategy. Evidence of antigenic drift is already evident in the most recent outbreaks of H5N1 (48).
    Several groups have developed reverse genetics?based methods that could speed the production of seed viruses as well as proposals for growing viruses in cell culture rather than in embyronated chicken eggs, which would allow for a much faster scale up in response to an epidemic (50). These technologies have not been approved yet for human use, though trials are underway.
    Even if the development of recombinant seed strains by reverse genetics becomes standard over the next few years, questions remain about how effective the current formaldehyde-inactivated seed strain strategies would be against pandemic strains, particularly the currently circulating H5N1 strains. Assuming that seed strains could be produced rapidly, several weeks would be required to manufacture a relevant number of doses of vaccine. To address this concern, several governments have been stockpiling vaccines based on H5N1 viruses that have been circulating over the last few years. While these vaccines may provide some protection, substantial evolution and antigenic drift seem to be occurring, rendering the stockpiled strains less and less useful (48).
    An approach based on conserved cellular epitopes within the internal genes has the advantage of subverting all of these issues. While cellular immunity is not sterilizing, it prevents illness and death in animal models (3). Common and immunodominant epitopes among circulating nonavian strains have been identified, and many of the same models and algorithms can be used to make predictions against the pathogenic strains (51). Mouse models are now available that have human leukocyte antigen (HLA) alleles, and they appear to recapitulate human epitope use. As described earlier, protection against death from highly pathogenic viruses has been shown in multiple systems. Cross-protective cell-mediated immunity has been found in birds for circulating chicken H5N1 and H9N2, both of which have been suggested as potential human pandemic strains (52). The notion of a "universal" vaccine for highly pathogenic strains is attractive.
    Antigenic drift due to immunologic pressure is also a concern with a CD8- or CD4-based vaccine approach. Reports have suggested that CD8+ epitopes under pressure will mutate to escape protective immunity (11). The mutation of an NP epitope that binds HLA-B35 present in strains of virus from the 1930s through the present indicates that even in nonpandemic years, immunologic pressure from cross-protective CD8+ T cells is enough to drive the evolution of the virus (53). In contrast, though, other dominant epitopes do not appear to be under the same pressure (54).
    Several human peptide epitopes that have been described and characterized show evidence of remarkably little mutation over many generations of viral evolution. In the most recent outbreaks of H5N1 virus, some of these peptides are conserved in viruses isolated from human patients (Table). The conservation of so many peptides from such distantly related viruses suggests that they may be less susceptible to antigenic drift than the HA and NA glycoproteins. Vaccines that promote strong memory CTL activity toward these peptides and MHC, in combination with the antibody-based approaches already underway, could help prevent pandemic influenza. This approach could potentiate immunologic pressure on the vaccine-targeted epitopes, but an immunization strategy that targets a large number of epitopes along with the natural restriction on epitope structure due to viral function should mitigate this effect. Some evidence shows that highly conserved CTL epitopes are restricted from mutation by viral structural requirements. Given the large number of influenza viruses sequenced over time, we should be able to make reasonable assumptions about the identity of these epitopes in MHC-diverse populations and focus on how to facilitate the development of strong immune responses toward them.
    This work was supported by USPHS grants AI29579 and CA21765 (to P.C.D.), AI065097 (to P.G.T.), and by ALSAC at St. Jude Children's Research Hospital. P.C.D. is a Burnet Fellow of the Australian National Health and Medical Research Council.
    Dr Thomas works in the Department of Immunology at St. Jude Children's Research Hospital in Memphis, Tennessee. His primary research involves the use of reverse genetics?engineered influenza viruses to understand the dynamics of CD8+ and CD4+ T-cell responses.
    References<sup>1</sup>
    1. Murphy BR, Webster RG. Orthomyxoviruses. In: Fields BN, Knipe DM, Howley PM, Chanock RM, Melnick JL, Monath TP, et al., editors. Fields virology. 3rd ed. Philadelphia: Lippincott-Raven Publishers; 1996. p. 1397?445.
    2. Allan W, Tabi Z, Cleary A, Doherty PC. Cellular events in the lymph node and lung of mice with influenza. Consequences of depleting CD4+ T cells. J Immunol. 1990;144:3980?6.
    3. Bender BS, Croghan T, Zhang L, Small PA Jr. Transgenic mice lacking class I major histocompatibility complex-restricted T cells have delayed viral clearance and increased mortality after influenza virus challenge. J Exp Med. 1992;175:1143?5.
    4. Hou S, Doherty PC, Zijlstra M, Jaenisch R, Katz JM. Delayed clearance of Sendai virus in mice lacking class I MHC-restricted CD8+ T cells. J Immunol. 1992;149:1319?25.
    5. Graham MB, Braciale TJ. Resistance to and recovery from lethal influenza virus infection in B lymphocyte-deficient mice. J Exp Med. 1997;186:2063?8.
    6. Epstein SL, Lo CY, Misplon JA, Bennink JR. Mechanism of protective immunity against influenza virus infection in mice without antibodies. J Immunol. 1998;160:322?7.
    7. Taylor PM, Askonas BA. Influenza nucleoprotein-specific cytotoxic T-cell clones are protective in vivo. Immunology. 1986;58:417?20.
    8. Andrew ME, Coupar BE. Efficacy of influenza haemagglutinin and nucleoprotein as protective antigens against influenza virus infection in mice. Scand J Immunol. 1988;28:81?5.
    9. Tite JP, Hughes-Jenkins C, O'Callaghan D, Dougan G, Russell SM, Gao XM, et al. Anti-viral immunity induced by recombinant nucleoprotein of influenza A virus. II. Protection from influenza infection and mechanism of protection. Immunology. 1990;71:202?7.
    10. Webster RG, Kawaoka Y, Taylor J, Weinberg R, Paoletti E. Efficacy of nucleoprotein and haemagglutinin antigens expressed in fowlpox virus as vaccine for influenza in chickens. Vaccine. 1991;9:303?8.
    11. Moskophidis D, Kioussis D. Contribution of virus-specific CD8+ cytotoxic T cells to virus clearance or pathologic manifestations of influenza virus infection in a T cell receptor transgenic mouse model. J Exp Med. 1998;188:223?32.
    12. Lawrence CW, Braciale TJ. Activation, differentiation, and migration of naive virus-specific CD8+ T cells during pulmonary influenza virus infection. J Immunol. 2004;173:1209?18.
    13. Tripp RA, Sarawar SR, Doherty PC. Characteristics of the influenza virus-specific CD8+ T cell response in mice homozygous for disruption of the H-2lAb gene. J Immunol. 1995;155:2955?9.
    14. Cerwenka A, Morgan TM, Dutton RW. Naive, effector, and memory CD8 T cells in protection against pulmonary influenza virus infection: homing properties rather than initial frequencies are crucial. J Immunol. 1999;163:5535?43.
    15. Lawrence CW, Ream RM, Braciale TJ. Frequency, specificity, and sites of expansion of CD8+ T cells during primary pulmonary influenza virus infection. J Immunol. 2005;174:5332?40.
    16. Walker JA, Molloy SS, Thomas G, Sakaguchi T, Yoshida T, Chambers TM, et al. Sequence specificity of furin, a proprotein-processing endoprotease, for the hemagglutinin of a virulent avian influenza virus. J Virol. 1994;68:1213?8.
    17. Walker JA, Sakaguchi T, Matsuda Y, Yoshida T, Kawaoka Y. Location and character of the cellular enzyme that cleaves the hemagglutinin of a virulent avian influenza virus. Virology. 1992;190:278?87.
    18. Prendergast JA, Helgason CD, Bleackley RC. A comparison of the flanking regions of the mouse cytotoxic cell proteinase genes. Biochim Biophys Acta. 1992;1131:192?8.
    19. Topham DJ, Tripp RA, Doherty PC. CD8+ T cells clear influenza virus by perforin or Fas-dependent processes. J Immunol. 1997;159:5197?200.
    20. Townsend AR, Rothbard J, Gotch FM, Bahadur G, Wraith D, McMichael AJ. The epitopes of influenza nucleoprotein recognized by cytotoxic T lymphocytes can be defined with short synthetic peptides. Cell. 1986;44:959?68.
    21. Belz GT, Xie W, Altman JD, Doherty PC. A previously unrecognized H-2D(b)-restricted peptide prominent in the primary influenza A virus-specific CD8(+) T-cell response is much less apparent following secondary challenge. J Virol. 2000;74:3486?93.
    22. Chen W, Calvo PA, Malide D, Gibbs J, Schubert U, Bacik I, et al. A novel influenza A virus mitochondrial protein that induces cell death. Nat Med. 2001;7:1306?12.
    23. Marshall DR, Turner SJ, Belz GT, Wingo S, Andreansky S, Sangster MY, et al. Measuring the diaspora for virus-specific CD8+ T cells. Proc Natl Acad Sci U S A. 2001;98:6313?8.
    24. McMichael AJ, Michie CA, Gotch FM, Smith GL, Moss B. Recognition of influenza A virus nucleoprotein by human cytotoxic T lymphocytes. J Gen Virol. 1986;67:719?26.
    25. Crowe SR, Miller SC, Brown DM, Adams PS, Dutton RW, Harmsen AG, et al. Uneven distribution of MHC class II epitopes within the influenza virus [corrected proof]. Vaccine. Epub 2005 Aug 15.
    26. Belz GT, Wodarz D, Diaz G, Nowak MA, Doherty PC. Compromised influenza virus-specific CD8(+)-T-cell memory in CD4(+)-T-cell-deficient mice. J Virol. 2002;76:12388?93.
    27. Riberdy JM, Christensen JP, Branum K, Doherty PC. Diminished primary and secondary influenza virus-specific CD8(+) T-cell responses in CD4-depleted Ig(-/-) mice. J Virol. 2000;74:9762?5.
    28. Brown DM, Roman E, Swain SL. CD4 T cell responses to influenza infection. Semin Immunol. 2004;16:171?7.
    29. Belz GT, Liu H, Andreansky S, Doherty PC, Stevenson PG. Absence of a functional defect in CD8+ T cells during primary murine gammaherpesvirus-68 infection of I-A(b-/-) mice. J Gen Virol. 2003;84:337?41.
    30. Diebold SS, Kaisho T, Hemmi H, Akira S, Reis e Sousa C. Innate antiviral responses by means of TLR7-mediated recognition of single-stranded RNA. Science. 2004;303:1529?31.
    31. Sarawar SR, Lee BJ, Reiter SK, Schoenberger SP. Stimulation via CD40 can substitute for CD4 T cell function in preventing reactivation of a latent herpesvirus. Proc Natl Acad Sci U S A. 2001;98:6325?9.
    32. Cardin RD, Brooks JW, Sarawar SR, Doherty PC. Progressive loss of CD8+ T cell-mediated control of a gamma-herpesvirus in the absence of CD4+ T cells. J Exp Med. 1996;184:863?71.
    33. Brooks JW, Hamilton-Easton AM, Christensen JP, Cardin RD, Hardy CL, Doherty PC. Requirement for CD40 ligand, CD4(+) T cells, and B cells in an infectious mononucleosis-like syndrome. J Virol. 1999;73:9650?4.
    34. Sun JC, Bevan MJ. Defective CD8 T cell memory following acute infection without CD4 T cell help. Science. 2003;300:339?42.
    35. Christensen JP, Cardin RD, Branum KC, Doherty PC. CD4(+) T cell-mediated control of a gamma-herpesvirus in B cell-deficient mice is mediated by IFN-gamma. Proc Natl Acad Sci U S A. 1999;96:5135?40.
    36. Appay V, Zaunders JJ, Papagno L, Sutton J, Jaramillo A, Waters A, et al. Characterization of CD4(+) CTLs ex vivo. J Immunol. 2002;168:5954?8.
    37. Norris PJ, Moffett HF, Yang OO, Kaufmann DE, Clark MJ, Addo MM, et al. Beyond help: direct effector functions of human immunodeficiency virus type 1-specific CD4(+) T cells. J Virol. 2004;78:8844?51.
    38. Staska LM, McGuire TC, Davies CJ, Lewin HA, Baszler TV. Neospora caninum?infected cattle develop parasite-specific CD4+ cytotoxic T lymphocytes. Infect Immun. 2003;71:3272?9.
    39. Roman E, Miller E, Harmsen A, Wiley J, von Andrian UH, Huston G, et al. CD4 effector T cell subsets in the response to influenza: heterogeneity, migration, and function. J Exp Med. 2002;196:957?68.
    40. Walker WS, Castrucci MR, Sangster MY, Carson RT, Kawaoka Y. HEL-Flu: an influenza virus containing the hen egg lysozyme epitope recognized by CD4+ T cells from mice transgenic for an alphabeta TCR. J Immunol. 1997;159:2563?6.
    41. Chapman TJ, Castrucci MR, Padrick RC, Bradley LM, Topham DJ. Antigen-specific and non-specific CD4(+) T cell recruitment and proliferation during influenza infection. Virology. 2005;340:296?306.
    42. Cole GA, Katz JM, Hogg TL, Ryan KW, Portner A, Woodland DL. Analysis of the primary T-cell response to Sendai virus infection in C57BL/6 mice: CD4+ T-cell recognition is directed predominantly to the hemagglutinin-neuraminidase glycoprotein. J Virol. 1994;68:6863?70.
    43. Horimoto T, Kawaoka Y. Influenza: lessons from past pandemics, warnings from current incidents. Nat Rev Microbiol. 2005;3:591?600.
    44. Riberdy JM, Flynn KJ, Stech J, Webster RG, Altman JD, Doherty PC. Protection against a lethal avian influenza A virus in a mammalian system. J Virol. 1999;73:1453?9.
    45. Christensen JP, Doherty PC, Branum KC, Riberdy JM. Profound protection against respiratory challenge with a lethal H7N7 influenza A virus by increasing the magnitude of CD8(+) T-cell memory. J Virol. 2000;74:11690?6.
    46. O'Neill E, Krauss SL, Riberdy JM, Webster RG, Woodland DL. Heterologous protection against lethal A/HongKong/156/97 (H5N1) influenza virus infection in C57BL/6 mice. J Gen Virol. 2000;81:2689?96.
    47. Cox RJ, Brokstad KA, Ogra P. Influenza virus: immunity and vaccination strategies. Comparison of the immune response to inactivated and live, attenuated influenza vaccines. Scand J Immunol. 2004;59:1?15.
    48. The World Health Organization Global Influenza Project Surveillance Network: Evolution of H5N1 avian influenza viruses in Asia. Emerg Infect Dis. 2005;11:1515?21.
    49. Jin H, Zhou H, Liu H, Chan W, Adhikary L, Mahmood K, et al. Two residues in the hemagglutinin of A/Fujian/411/02-like influenza viruses are responsible for antigenic drift from A/Panama/2007/99. Virology. 2005;336:113?9.
    50. Ozaki H, Govorkova EA, Li C, Xiong X, Webster RG, Webby RJ. Generation of high-yielding influenza A viruses in African green monkey kidney (Vero) cells by reverse genetics. J Virol. 2004;78:1851?7.
    51. Cheuk E, Chamberlain JW. Strong memory CD8(+) T cell responses against immunodominant and three new subdominant HLA-B27-restricted influenza A CTL epitopes following secondary infection of HLA-B27 transgenic mice. Cell Immunol. 2005;234:110?23.
    52. Seo SH, Peiris M, Webster RG. Protective cross-reactive cellular immunity to lethal A/Goose/Guangdong/1/96-like H5N1 influenza virus is correlated with the proportion of pulmonary CD8(+) T cells expressing gamma interferon. J Virol. 2002;76:4886?90.
    53. Boon AC, de Mutsert G, Graus YM, Fouchier RA, Sintnicolaas K, Osterhaus AD, et al. Sequence variation in a newly identified HLA-B35-restricted epitope in the influenza A virus nucleoprotein associated with escape from cytotoxic T lymphocytes. J Virol. 2002;76:2567?72.
    54. Berkhoff EG, de Wit E, Geelhoed-Mieras MM, Boon AC, Symons J, Fouchier RA, et al. Functional constraints of influenza a virus epitopes limit escape from cytotoxic T lymphocytes. J Virol. 2005;79:11239?46.
    55. Macken C, Lu H, Goodman J, Boykin L. The value of a database in surveillance and vaccine selection. In: Osterhaus ADME, Cox N, Hampson AW, editors. Options for the control of influenza IV. Amsterdam: Elsevier Science; 2001. p. 103?6.

    <sup>1</sup>Further literature support for the material discussed in this article is available in the Appendix Bibliography.

    <table border="0" cellpadding="0" cellspacing="0" width="100%"> <tbody><tr> <td colspan="4" valign="bottom"> Table. Conservation of human NP and M1 epitopes between H1N1 PR8 and 3 human isolates of H5N1 viruses (A/Hong Kong/156/1997, A/Hong Kong/213/2003, and A/Vietnam/1203/2004)*
    </td> </tr> <tr> <td colspan="4"> <hr noshade="noshade" size="1"> </td> </tr> <tr> <td valign="bottom"> Epitope
    </td> <td valign="bottom">
    HLA restriction
    </td> <td valign="bottom">
    PR8 sequence
    </td> <td valign="bottom">
    Conservation
    </td> </tr> <tr> <td colspan="4"> <hr noshade="noshade" size="1"> </td> </tr> <tr> <td valign="top"> NP 383?391
    </td> <td valign="top">
    B*2705
    </td> <td valign="top">
    SRYWAIRTR
    </td> <td valign="top">
    3/3 identical
    </td> </tr> <tr> <td valign="top"> NP 418?426
    </td> <td valign="top">
    B*3501
    </td> <td valign="top">
    LPFDRTTIM
    </td> <td valign="top">
    0/3 identical
    </td> </tr> <tr> <td valign="top"> NP 44?52
    </td> <td valign="top">
    A*01
    </td> <td valign="top">
    CTELKLSDY
    </td> <td valign="top">
    2/3 identical (156 Y9Q)
    </td> </tr> <tr> <td valign="top"> NP 265?273
    </td> <td valign="top">
    A*03
    </td> <td valign="top">
    ILRGSVAHK
    </td> <td valign="top">
    3/3 identical
    </td> </tr> <tr> <td valign="top"> NP 188?198
    </td> <td valign="top">
    A*1101
    </td> <td valign="top">
    TMVMELVRMIK
    </td> <td valign="top">
    3/3 V7I mutation
    </td> </tr> <tr> <td valign="top"> NP 380?388
    </td> <td valign="top">
    B*08
    </td> <td valign="top">
    ELRSRYWAI
    </td> <td valign="top">
    3/3 identical
    </td> </tr> <tr> <td valign="top"> NP 174?184
    </td> <td valign="top">
    B*2705
    </td> <td valign="top">
    RRSGAAGAAVK
    </td> <td valign="top">
    2/3 identical (156 V10I)
    </td> </tr> <tr> <td valign="top"> M1 58?66
    </td> <td valign="top">
    A*0201
    </td> <td valign="top">
    GILGFVFTL
    </td> <td valign="top">
    3/3 identical
    </td> </tr> <tr> <td valign="top"> M1 27?35
    </td> <td valign="top">
    A*03
    </td> <td valign="top">
    RLEDVFAGK
    </td> <td valign="top">
    2/3 mutated (1203, 213 both R1K)
    </td> </tr> <tr> <td valign="top"> M1 13?21
    </td> <td valign="top">
    A*1101
    </td> <td valign="top">
    SIIPSGPLK
    </td> <td valign="top">
    3/3 identical
    </td> </tr> <tr> <td colspan="4"> <hr noshade="noshade" size="1"> </td> </tr> <tr> <td colspan="4" valign="top"> *All 3 isolates were compared to the mouse-adapted PR8 strain and differences are reported. Sequences obtained from the Influenza Sequence Database (55). NP, nucleoprotein; HLA, human leukocyte antigen.
    </td> </tr> </tbody></table>

  • #2
    Re: Cell-mediated Protection in Influenza Infection

    Functional Constraints of Influenza A Virus Epitopes Limit Escape from Cytotoxic T Lymphocytes


    E. G. M. Berkhoff, E. de Wit, M. M. Geelhoed-Mieras, A. C. M. Boon, J. Symons, R. A. M. Fouchier, A. D. M. E. Osterhaus, and G. F. Rimmelzwaan<sup>*</sup> Department of Virology and WHO National Influenza Center, Erasmus Medical Center, Rotterdam, The Netherlands
    Received 16 March 2005/ Accepted 9 June 2005
    <!-- null -->
    <table bgcolor="#e1e1e1" cellpadding="0" cellspacing="0" width="100%"> <tbody><tr><td align="left" bgcolor="#ffffff" valign="middle" width="5%"></td> <th align="left" valign="middle" width="95%"> ABSTRACT </th></tr></tbody></table> <table align="right" border="1" cellpadding="5"><tbody><tr><th align="left"> Top
    Abstract
    Introduction
    Materials and Methods
    Results
    Discussion
    References
    </th></tr></tbody></table>
    Viruses can exploit a variety of strategies to evade immune<sup> </sup>surveillance by cytotoxic T lymphocytes (CTL), including the<sup> </sup>acquisition of mutations in CTL epitopes. Also for influenza<sup> </sup>A viruses a number of amino acid substitutions in the nucleoprotein<sup> </sup>(NP) have been associated with escape from CTL. However, other<sup> </sup>previously identified influenza A virus CTL epitopes are highly<sup> </sup>conserved, including the immunodominant HLA-A*0201-restricted<sup> </sup>epitope from the matrix protein, M1<sub>58-66</sub>. We hypothesized that<sup> </sup>functional constraints were responsible for the conserved nature<sup> </sup>of influenza A virus CTL epitopes, limiting escape from CTL.<sup> </sup>To assess the impact of amino acid substitutions in conserved<sup> </sup>epitopes on viral fitness and recognition by specific CTL, we<sup> </sup>performed a mutational analysis of CTL epitopes. Both alanine<sup> </sup>replacements and more conservative substitutions were introduced<sup> </sup>at various positions of different influenza A virus CTL epitopes.<sup> </sup>Alanine replacements for each of the nine amino acids of the<sup> </sup>M1<sub>58-66</sub> epitope were tolerated to various extents, except for<sup> </sup>the anchor residue at the second position. Substitution of anchor<sup> </sup>residues in other influenza A virus CTL epitopes also affected<sup> </sup>viral fitness. Viable mutant viruses were used in CTL recognition<sup> </sup>experiments. The results are discussed in the light of the possibility<sup> </sup>of influenza viruses to escape from specific CTL. It was speculated<sup> </sup>that functional constraints limit variation in certain epitopes,<sup> </sup>especially at anchor residues, explaining the conserved nature<sup> </sup>of these epitopes.<sup> </sup>
    <!-- null -->
    <table bgcolor="#e1e1e1" cellpadding="0" cellspacing="0" width="100%"> <tbody><tr><td align="left" bgcolor="#ffffff" valign="middle" width="5%"></td> <th align="left" valign="middle" width="95%"> INTRODUCTION </th></tr></tbody></table> <table align="right" border="1" cellpadding="5"><tbody><tr><th align="left"> Top
    Abstract
    Introduction
    Materials and Methods
    Results
    Discussion
    References
    </th></tr></tbody></table>
    Cytotoxic T lymphocytes (CTL) play an important role in the<sup> </sup>control of viral infections (10). To evade these CTL responses,<sup> </sup>viruses can exploit a variety of mechanisms to prevent recognition<sup> </sup>by specific CTL, including the accumulation of amino acid substitutions<sup> </sup>in or adjacent to CTL epitopes (28, 39). This strategy has been<sup> </sup>shown predominantly by certain RNA viruses, such as human immunodeficiency<sup> </sup>virus (HIV) (7, 21, 22, 35, 44, 47, 48), hepatitis C virus (9,<sup> </sup>60), and lymphocytic choriomeningitis virus (36, 46), which<sup> </sup>are known for their high mutation rate. Also for influenza A<sup> </sup>viruses, a number of amino acid substitutions in the nucleoprotein<sup> </sup>(NP) have been associated with escape from human CTL. One of<sup> </sup>them, the R-to-G substitution at position 384 (R384G), which<sup> </sup>is at the anchor residues of the HLA-B*0801-restricted NP<sub>380-388</sub><sup> </sup>and HLA-B*2705-restricted NP<sub>383-391</sub> epitopes, resulted in the<sup> </sup>loss of these epitopes (51, 58). This substitution reduced the<sup> </sup>in vitro virus-specific CTL response in HLA-B*2705-positive<sup> </sup>individuals significantly (2). Although the R384G substitution<sup> </sup>was tolerated only in the presence of one or more functionally<sup> </sup>compensating comutations (50, 52), it was fixed rapidly. This<sup> </sup>was explained by small selective advantages and population dynamics<sup> </sup>in a theoretical model (19). A third variable epitope in the<sup> </sup>influenza A virus NP is the HLA-B*3501-restricted epitope NP<sub>418-426</sub>,<sup> </sup>which displayed considerable variability in T-cell contact residues,<sup> </sup>affecting recognition by specific CTL (4, 5). Thus, in contrast<sup> </sup>to the two epitopes described above, the NP<sub>418-426</sub> epitope retained<sup> </sup>its anchor residues for binding to HLA-B*3501. Other previously<sup> </sup>identified epitopes in influenza A virus proteins are highly<sup> </sup>conserved, such as the immunodominant HLA-A*0201-restricted<sup> </sup>epitope from the matrix protein, M1<sub>58-66</sub>. It is likely that<sup> </sup>selective pressure by CTL against this epitope is high, considering<sup> </sup>the immunodominant nature of the epitope (3, 55) and the high<sup> </sup>prevalence of HLA-A*0201 in the human population (34). Yet,<sup> </sup>the amino acid sequence of this nine-mer epitope is conserved,<sup> </sup>even between different subtypes of human influenza A virus.<sup> </sup>We hypothesize that functional constraints are responsible for<sup> </sup>the inability of the virus to accumulate amino acid substitutions<sup> </sup>in this and other conserved epitopes, limiting immune escape<sup> </sup>from virus-specific CTL responses. To test this hypothesis,<sup> </sup>we performed a mutational analysis of various epitopes and tested<sup> </sup>the effect of selected amino acid substitutions on viral fitness<sup> </sup>and immune recognition by CTL. For this purpose, we employed<sup> </sup>a plasmid-driven rescue system for the generation of recombinant<sup> </sup>influenza viruses. For the epitope M1<sub>58-66</sub> (GILGFVFTL), we performed<sup> </sup>alanine replacements for each of the nine amino acid positions.<sup> </sup>In addition, various other amino acid substitutions were introduced<sup> </sup>in this and four other epitopes, namely the HLA-A*0101-restricted<sup> </sup>epitopes PB1<sub>591-599</sub> and NP<sub>44-52</sub>, the HLA-B*2705-restricted epitope<sup> </sup>NP<sub>174-184</sub>, and the HLA-B*3501-restricted epitope NP<sub>418-426</sub>.<sup> </sup>Single mutations at anchor residues could result in the loss<sup> </sup>of the epitopes, which would constitute the most economical<sup> </sup>way for the virus to escape from immune surveillance by specific<sup> </sup>CTL. Therefore, we focused on the mutational analysis of anchor<sup> </sup>residues of the respective epitopes. The data obtained in the<sup> </sup>present study on viral fitness and recognition of influenza<sup> </sup>viruses with mutations in CTL epitopes are discussed in the<sup> </sup>light of natural evolution of CTL epitopes.<sup> </sup>
    <!-- null -->
    <table bgcolor="#e1e1e1" cellpadding="0" cellspacing="0" width="100%"> <tbody><tr><td align="left" bgcolor="#ffffff" valign="middle" width="5%"></td> <th align="left" valign="middle" width="95%"> MATERIALS AND METHODS </th></tr></tbody></table> <table align="right" border="1" cellpadding="5"><tbody><tr><th align="left"> Top
    Abstract
    Introduction
    Materials and Methods
    Results
    Discussion
    References
    </th></tr></tbody></table>
    Plasmids. For the generation of recombinant influenza viruses, a bidirectional<sup> </sup>reverse genetics system based on influenza virus A/Netherlands/178/95<sup> </sup>(A/NL/178/95; H3N2), was used. The viral gene segments were<sup> </sup>amplified by reverse transcription-PCR using segment-specific<sup> </sup>primers, purified by electrophoresis in agarose gels according<sup> </sup>to standard methods, and cloned into a modified pHW2000 vector<sup> </sup>as previously described (11, 24). Subsequently, site-directed<sup> </sup>mutagenesis was performed (QuikChange site-directed mutagenesis<sup> </sup>kit; Stratagene, La Jolla, CA) to substitute single amino acids<sup> </sup>in several influenza virus CTL epitopes, as listed in Table<sup> </sup>1. Sequence analysis was performed for all recombinant plasmids,<sup> </sup>using a Big Dye Terminator v3.1 cycle sequencing kit (Applied<sup> </sup>Biosystems, Foster City, CA) and an ABI PRISM 3100 genetic analyzer<sup> </sup>(Applied Biosystems), according to the instructions of the manufacturer.<sup> </sup>All PCR primer sequences and plasmid maps are available on request.<sup> </sup>
    <!-- null -->

    <center><table cellpadding="0" cellspacing="0" width="95%"><tbody><tr bgcolor="#e1e1e1"><td><table cellpadding="2" cellspacing="2"> <tbody><tr bgcolor="#e1e1e1"><td align="center" bgcolor="#ffffff" valign="top"> View this table:
    <nobr>[in this window]
    [in a new window]
    </nobr> </td><td align="left" valign="top"> TABLE 1. Substitutions introduced in CTL epitopes in this study
    </td></tr></tbody></table> </td></tr></tbody></table></center>
    Generation of viruses. The recombinant bidirectional plasmids were transfected into<sup> </sup>293T cells, using the calcium phosphate precipitation method<sup> </sup>as described previously (11). After 48 h, culture supernatants<sup> </sup>were harvested and used for subsequent infection of confluent<sup> </sup>Madin-Darby canine kidney (MDCK) cells (2). After 3 days, culture<sup> </sup>supernatants were harvested, cleared by low-speed centrifugation,<sup> </sup>aliquoted, and stored at ?80?C until use. The recombinant<sup> </sup>viruses were designated "A/NL/95-" followed by their specific<sup> </sup>substitutions. In order to confirm the introduction of the mutations<sup> </sup>and to exclude the introduction of second site mutations, the<sup> </sup>nucleotide sequences of the corresponding full-length genes<sup> </sup>were assessed. Infectious virus titers were determined as previously<sup> </sup>described (49). Multistep growth kinetics were determined for<sup> </sup>all recombinant influenza viruses after infection of MDCK cells,<sup> </sup>using an equivalent multiplicity of infection (MOI) of 0.001<sup> </sup>50% tissue culture infectious dose (TCID<sub>50</sub>) per cell, which<sup> </sup>was used as a measure of viral fitness. Viral fitness was considered<sup> </sup>reduced when a statistically significant difference was observed<sup> </sup>compared to plasmid-derived wild-type virus.<sup> </sup> CD8<sup>+</sup>-T-cell clones. Generation of CD8<sup>+</sup>-T-cell clones directed against the HLA-A*0201-restricted<sup> </sup>epitope M1<sub>58-66</sub>, the HLA-B*3501-restricted epitope NP<sub>418-426</sub>,<sup> </sup>and the HLA-A*0101-restricted epitopes NP<sub>44-52</sub> and PB1<sub>591-599</sub><sup> </sup>was described previously (5, 58).<sup> </sup>
    Target cells. B-lymphoblastoid cell lines, established as described previously<sup> </sup>(53), and three C1R cell lines, kindly provided by P. Romero<sup> </sup>(HLA-A*0201-transfected C1R cell line), M. Takiguchi (HLA-B*3501-transfected<sup> </sup>C1R cell line), and P. Cresswell (HLA-A*0101-transfected C1R<sup> </sup>cell line), were used as target cells. Peptide labeling was<sup> </sup>performed by incubating 10<sup>6</sup> cells/ml overnight with 5 ?M<sup> </sup>peptide in RPMI 1640 medium (Cambrex, East Rutherford, NJ) supplemented<sup> </sup>with 10% fetal calf serum (FCS) and antibiotics (R10F). Peptides<sup> </sup>were manufactured, high-performance liquid chromatography purified<sup> </sup>(immunograde, >85% purity), and analyzed by mass spectrometry<sup> </sup>(Eurogentec, Seraing, Belgium). For infection with the recombinant<sup> </sup>influenza viruses, 10<sup>6</sup> target cells were infected at an MOI<sup> </sup>of 3 in a volume of 1 ml. After incubation for 1 h at 37?C,<sup> </sup>the cells were resuspended in R10F and incubated for 16 to 18<sup> </sup>h.<sup> </sup>
    Intracellular IFN- staining and flow cytometry. The CD8<sup>+</sup>-T-cell clones were adjusted to a concentration of 10<sup>6</sup><sup> </sup>cells/ml in R10F supplemented with Golgistop (monensin; Pharmingen,<sup> </sup>Alphen a/d Rijn, The Netherlands). Sixty thousand effector cells<sup> </sup>were incubated with 3 x 10<sup>5</sup> stimulator cells, which were infected,<sup> </sup>pulsed with peptides, or left untreated, for 6 h at 37?C<sup> </sup>in U-bottom plates. Subsequently intracellular gamma interferon<sup> </sup>(IFN-) staining was performed as described previously (6). In<sup> </sup>brief, the cells were washed with phosphate-buffered saline<sup> </sup>(PBS) containing 2% FCS (P2F) and Golgistop, stained with monoclonal<sup> </sup>antibody (MAb) directed to CD3 (Becton Dickinson, Alphen a/d<sup> </sup>Rijn, The Netherlands) and CD8 (Dako, Glostrup, Denmark), fixed<sup> </sup>and permeabilized with Cytofix and Cytoperm (Pharmingen), and<sup> </sup>stained with a MAb specific for IFN- (Pharmingen). At least<sup> </sup>5 x 10<sup>3</sup> gated CD3<sup>+</sup> CD8<sup>+</sup> events were acquired using a FACSCalibur<sup> </sup>(Becton Dickinson) flow cytometer. The data were analyzed using<sup> </sup>the software program Cell Quest Pro (Becton Dickinson).<sup> </sup>
    ELISPOT. Enzyme-linked immunospot (ELISPOT) assays were performed as<sup> </sup>described previously (3). In brief, 96-well Silent Screen plates<sup> </sup>(Nalge Nunc International, Rochester, NY) were coated with 2.5<sup> </sup>?g/ml of anti-IFN- MAb 1-DIK (Mabtech, Stockholm, Sweden)<sup> </sup>and blocked with RPMI 1640 medium supplemented with 10% human<sup> </sup>AB serum (Sanquin Bloodbank, Rotterdam, The Netherlands), antibiotics,<sup> </sup>and 20 ?M ?-mercaptoethanol (R10H). Three thousand<sup> </sup>cells of CD8<sup>+</sup>-T-cell clones were incubated with 3 x 10<sup>4</sup> target<sup> </sup>cells, which were infected, pulsed with peptides, or left untreated,<sup> </sup>for 4 h. Next, the plates were washed with PBS-0.05% Tween 20<sup> </sup>(Sigma Chemical Co., St. Louis, MO), and secreted IFN- was detected<sup> </sup>using biotinylated anti-IFN- MAb 7-B6-1 (Mabtech; dilution of<sup> </sup>1:5,000). Subsequently, streptavidin labeled with alkaline phosphatase<sup> </sup>was added, which was visualized with the phosphatase substrate<sup> </sup>BCIP/NBT (5-bromo-4-chloro-3-indolylphosphate/nitroblue tetrazolium)<sup> </sup>(Kirkegaard & Perry Laboratories, Gaithersburg, MD). Numbers<sup> </sup>of spots were counted using an automated image analysis ELISPOT<sup> </sup>reader (Aelvis; Sanquin Bloodbank, Amsterdam, The Netherlands).<sup> </sup>
    Chromium release assay. Chromium release assays were performed as described previously<sup> </sup>(1). In brief, 7.5 x 10<sup>5</sup> cells per target were labeled with<sup> </sup>75 ?Ci Na<sub>2</sub>[<sup>51</sup>Cr]O<sub>4</sub> and incubated with CD8<sup>+</sup>-T-cell clone<sup> </sup>at effector/target (E:T) ratios of 10, 5, 2.5, and 1. Target<sup> </sup>cells were also incubated with 10% Triton X-100 or R10F to determine<sup> </sup>the maximum and spontaneous release, respectively. After 4 h<sup> </sup>of incubation, the supernatants were harvested (Skatron Instruments,<sup> </sup>Sterling, VA) and radioactivity was measured by gamma counting.<sup> </sup>The percentage of specific lysis was calculated by the following<sup> </sup>formula: [(experimental release ? spontaneous release)/(maximum<sup> </sup>release ? spontaneous release)] x 100%. The chromium release<sup> </sup>assays were performed in quadruplicate, and the data are presented<sup> </sup>as the average.<sup> </sup>
    Synonymous/nonsynonymous analysis. The ratio of synonymous and nonsynonymous nucleotide substitutions<sup> </sup>was calculated using a synonymous/nonsynonymous analysis program<sup> </sup>(SNAP) (27, 38, 40) at www.hiv.lanl.gov. The NP nucleotide sequences<sup> </sup>of influenza viruses A/England/878/69 (AY210221<!-- HIGHWIRE EXLINK_ID="79:17:11239:1" VALUE="AY210221" TYPEGUESS="GEN" --><!-- /HIGHWIRE -->) and A/New York/12/2003<sup> </sup>(CY000124<!-- HIGHWIRE EXLINK_ID="79:17:11239:2" VALUE="CY000124" TYPEGUESS="GEN" --><!-- /HIGHWIRE -->) obtained from the Influenza Sequence Database (www.flu.lanl.gov)<sup> </sup>(32) were used in comparison for analysis by SNAP. These viruses<sup> </sup>were selected since the NP genes belonged to the same lineage<sup> </sup>of influenza A (H3N2) viruses (30).<sup> </sup>
    <!-- null -->
    <table bgcolor="#e1e1e1" cellpadding="0" cellspacing="0" width="100%"> <tbody><tr><td align="left" bgcolor="#ffffff" valign="middle" width="5%"></td> <th align="left" valign="middle" width="95%"> RESULTS </th></tr></tbody></table> <table align="right" border="1" cellpadding="5"><tbody><tr><th align="left"> Top
    Abstract
    Introduction
    Materials and Methods
    Results
    Discussion
    References
    </th></tr></tbody></table>
    Synonymous/nonsynonymous analysis. In order to obtain an impression of the selective pressure mediated<sup> </sup>by virus-specific CTL, we performed a synonymous/nonsynonymous<sup> </sup>analysis with the NP nucleotide sequences of a pair of influenza<sup> </sup>A (H3N2) viruses consisting of influenza virus A/England/878/69,<sup> </sup>isolated shortly after the introduction of H3N2 virus in the<sup> </sup>human population, and a more recent strain, A/New York/12/2003.<sup> </sup>The NP gene was selected for this type of analysis since 14<sup> </sup>of the known epitopes are located within this protein. The synonymous/nonsynonymous<sup> </sup>(ds/dn) ratio for the sequence encoding the 14 epitopes was<sup> </sup>8.67, whereas the ds/dn ratio for the rest of the protein was<sup> </sup>19.73, which is suggestive for selective pressure on the CTL<sup> </sup>epitopes. However, BLAST search of up to 450 influenza A H3N2<sup> </sup>virus sequences indicated that all known CTL epitopes, including<sup> </sup>those located within other viral proteins, retained their anchor<sup> </sup>residues, with the exception of the HLA-B*0801- and HLA-B*2705-restricted<sup> </sup>epitopes NP<sub>380-388</sub> and NP<sub>383-391</sub> (see Discussion).<sup> </sup>
    Viral fitness of influenza viruses with mutations in CTL epitopes. Since the HLA-A*0201-restricted epitope M1<sub>58-66</sub> (GILGFVFTL)<sup> </sup>is highly conserved, we selected this epitope to examine the<sup> </sup>effect on viral fitness of alanine replacements at each of the<sup> </sup>nine positions of the epitope. Mutant viruses could be rescued<sup> </sup>with alanine replacements at all positions within the M1<sub>58-66</sub><sup> </sup>epitope, except for the second position (Fig. 1A). The alanine<sup> </sup>replacement at position 59 of the matrix protein, which is the<sup> </sup>anchor residue of the M1<sub>58-66</sub> epitope, was detrimental to viral<sup> </sup>fitness. Although viruses with alanine replacements at the other<sup> </sup>eight positions were rescued, the virus replication kinetics<sup> </sup>of these mutants was affected compared to that of wild-type<sup> </sup>virus (Fig. 1B). Especially, mutant viruses A/NL/95-M1 F62A<sup> </sup>and -M1 F64A yielded >100-fold less progeny virus than the<sup> </sup>wild-type virus at 12 h postinfection. At 48 h postinfection,<sup> </sup>the differences from wild-type virus were still at least 50-fold.<sup> </sup>In addition, the L60A substitution caused a reduction of 75-fold<sup> </sup>in virus production compared to wild-type virus from 24 h onwards.<sup> </sup>Since the dramatic effect of the alanine replacement at the<sup> </sup>anchor residue was of special interest, we decided to study<sup> </sup>the effect of more conservative substitutions at position 59.<sup> </sup>We also replaced the isoleucine at this position with a leucine<sup> </sup>and a valine (M1 I59L and I59V) and found that in contrast to<sup> </sup>M1 I59A these replacements were tolerated by the virus to a<sup> </sup>certain extent, since these mutant viruses were readily rescued<sup> </sup>(Fig. 1C). We also performed multistep growth curves with these<sup> </sup>mutant viruses. Six and 12 h postinfection, these mutant viruses<sup> </sup>yielded 100-fold and 30-fold less progeny virus than wild-type<sup> </sup>virus, respectively. From 24 h postinfection onwards, these<sup> </sup>differences were no longer statistically significant (Fig. 1D).<sup> </sup>
    <!-- null -->

    <center><table cellpadding="0" cellspacing="0" width="95%"><tbody><tr bgcolor="#e1e1e1"><td><table cellpadding="2" cellspacing="2"> <tbody><tr bgcolor="#e1e1e1"><td align="center" bgcolor="#ffffff" valign="top">
    View larger version (19K):
    <nobr>[in this window]
    [in a new window]
    </nobr> </td><td align="left" valign="top"> FIG. 1. Effect of amino acid substitutions in the M1<sub>58-66</sub> epitope on viral fitness. Upon transfection of 293T cells and subsequent rescue in MDCK cells, infectious virus titers were determined for the wild-type (WT) and mutant influenza viruses with alanine replacements for each of the nine amino acids of the M1<sub>58-66</sub> epitope (A) and with the more conservative substitutions at position 59 (C). Influenza virus could be rescued with alanine replacements at all positions within the M1<sub>58-66</sub> epitope, except for position 59 (A). Influenza A virus tolerated the more conservative substitutions M1 I59L and I59V to a certain extent (C). The data represent the average of three experiments. Subsequently, growth curves were generated (B and D) postinfection (p.i.) of MDCK cells at an MOI of 0.001. Virus replication kinetics for wild-type virus (?) and M1 G58A (), L60A (), G61A (), F62A (), V63A (), F64A (), T65A (), L66A () (B), and wild-type virus (?), I59L (), and I59V () (D) are shown. The data represent the average of three experiments. The error bars indicate the standard deviation. An asterisk indicates statistical significance (P < 0.05, Student's t test).
    </td></tr></tbody></table> </td></tr></tbody></table></center>
    Because of the impact of the alanine replacement at the anchor<sup> </sup>residue of the M1<sub>58-66</sub> epitope, we decided to study the effect<sup> </sup>on anchor residue substitutions of the HLA-B*3501-restricted<sup> </sup>epitope NP<sub>418-426</sub> (LPFEKSTVM). This epitope displays a high<sup> </sup>degree of variability but retained its anchor residues for binding<sup> </sup>to HLA-B*3501 (4). Replacement of the proline at position 419<sup> </sup>or the methionine at position 426 with an alanine at these positions<sup> </sup>was detrimental to viral fitness (Fig. 2A). A more conservative<sup> </sup>substitution at position 419 (NP P419G) also prevented rescue<sup> </sup>of viable virus. The conservative NP M426I substitution was<sup> </sup>not detrimental to viral fitness, and only 6-h postinfection<sup> </sup>virus replication kinetics were significantly impaired (38-fold)<sup> </sup>compared to those of wild-type virus (Fig. 2A and B).<sup> </sup> <!-- null -->

    <center><table cellpadding="0" cellspacing="0" width="95%"><tbody><tr bgcolor="#e1e1e1"><td><table cellpadding="2" cellspacing="2"> <tbody><tr bgcolor="#e1e1e1"><td align="center" bgcolor="#ffffff" valign="top">
    View larger version (20K):
    <nobr>[in this window]
    [in a new window]
    </nobr> </td><td align="left" valign="top"> FIG. 2. Effect of amino acid substitutions in the NP<sub>418-426</sub> epitope on viral fitness. Upon transfection of 293T cells and subsequent rescue in MDCK cells, infectious virus titers were determined for the wild-type and mutant influenza viruses with alanine replacements at position 419 or 426 or with the more conservative substitutions NP P419G and M426I (A). The data represent the average of three experiments. Subsequently, growth curves of wild-type (WT) virus (?) and influenza virus A/NL/95-NP M426I () were generated postinfection (p.i.) of MDCK cells at an MOI of 0.001 (B). The data represent the average of three experiments. The error bars indicate the standard deviation. An asterisk indicates statistical significance (P < 0.05, Student's t test).
    </td></tr></tbody></table> </td></tr></tbody></table></center>
    In addition, conservative amino acid substitutions at the anchor<sup> </sup>residues of the HLA-A*0101-restricted epitopes PB1<sub>591-599</sub> (VSDGGPNLY)<sup> </sup>and NP<sub>44-52</sub> (CTELKLSDY) and the HLA-B*2705-restricted epitope<sup> </sup>NP<sub>174-184</sub> (RRSGAAGAAVK) were introduced. The D593N substitution<sup> </sup>in PB1 was detrimental to viral fitness. The E46Q and R175K<sup> </sup>substitutions in the viral NP were not detrimental to viral<sup> </sup>fitness (Fig. 3A), although mutant virus A/NL/95-NP E46Q yielded<sup> </sup>up to 133-fold less progeny virus within the first 24 h postinfection<sup> </sup>than wild-type virus. For mutant virus A/NL/95-NP R175K, significantly<sup> </sup>lower virus titers were observed than for wild-type virus from<sup> </sup>48 h postinfection onwards (Fig. 3B).<sup> </sup> <!-- null -->

    <center><table cellpadding="0" cellspacing="0" width="95%"><tbody><tr bgcolor="#e1e1e1"><td><table cellpadding="2" cellspacing="2"> <tbody><tr bgcolor="#e1e1e1"><td align="center" bgcolor="#ffffff" valign="top">
    View larger version (21K):
    <nobr>[in this window]
    [in a new window]
    </nobr> </td><td align="left" valign="top"> FIG. 3. Effect of amino acid substitutions on viral fitness. Upon transfection of 293T cells and subsequent rescue in MDCK cells, infectious virus titers were determined for the wild-type (WT) and mutant influenza viruses with the conservative amino acid substitution PB1 D593N, NP E46Q, or NP R175K in epitopes PB1<sub>591-599</sub>, NP<sub>44-52</sub>, and NP<sub>174-184</sub>, respectively (A). The data represent the average of three experiments. Subsequently, growth curves of wild-type virus (?) and mutant viruses A/NL/95-NP E46Q () and R175K () were generated postinfection (p.i.) of MDCK cells at an MOI of 0.001 (B). The data represent the average of three experiments. The error bars indicate the standard deviation. An asterisk indicates statistical significance between the wild type and mutant influenza virus A/NL/95-NP E46Q at 6 and 12 h postinfection and between the wild type and influenza virus A/NL/95-NP R175K at 48 and 72 h postinfection (P < 0.05, Student's t test).
    </td></tr></tbody></table> </td></tr></tbody></table></center>
    Recognition of mutant viruses by CTL. The recognition of HLA-A*0201-positive cells infected with the<sup> </sup>various M1<sub>58-66</sub> mutant viruses by specific CTL was determined<sup> </sup>by intracellular IFN- staining, ELISPOT assay, and classical<sup> </sup>chromium release assays. As shown in Fig. 4, the M1<sub>58-66</sub>-specific<sup> </sup>CTL clone recognized C1R-A2 cells infected with wild-type virus<sup> </sup>and mutant A/NL/95 virus with the M1 G58A, L60A, or L66A substitution,<sup> </sup>but not cells infected with A/NL/95 mutant viruses with the<sup> </sup>M1 G61A, F62A, V63A, or T65A substitution or noninfected cells,<sup> </sup>as determined by intracellular IFN- staining and flow cytometry.<sup> </sup>These observations were confirmed by ELISPOT (Fig. 4K and L)<sup> </sup>and by chromium release assays (Fig. 4M and N). A control CTL<sup> </sup>clone specific for the NP<sub>418-426</sub> epitope recognized target cells<sup> </sup>infected with all mutant A/NL/95 viruses similarly, indicating<sup> </sup>that the infection of the cells and the processing and presentation<sup> </sup>of immunogenic peptides were comparable for all viruses (Fig.<sup> </sup>4L and N). The recognition of A/NL/95-M1 I59A and F64A could<sup> </sup>not be tested, since these mutant viruses could not be propagated<sup> </sup>to sufficiently high titers. C1R-A2 cells infected with A/NL/95<sup> </sup>mutant viruses with the more conservative amino acid substitutions<sup> </sup>M1 I59V and I59L were fully recognized in all three assays (Fig.<sup> </sup>5).<sup> </sup> <!-- null -->

    <center><table cellpadding="0" cellspacing="0" width="95%"><tbody><tr bgcolor="#e1e1e1"><td><table cellpadding="2" cellspacing="2"> <tbody><tr bgcolor="#e1e1e1"><td align="center" bgcolor="#ffffff" valign="top">
    View larger version (42K):
    <nobr>[in this window]
    [in a new window]
    </nobr> </td><td align="left" valign="top"> FIG. 4. Effect of alanine replacements in the M1<sub>58-66</sub> epitope on recognition by specific CTL. Reactivity of CTL clone, directed against the M1<sub>58-66</sub> epitope, with stimulator cells infected with wild type (WT) (B) or influenza virus A/NL/95-M1 G58A (C), L60A (D), G61A (E), F62A (F), V63A (G), T65A (H), or L66A (I) was determined by intracellular IFN- staining and flow cytometry. PE, phycoerythrin conjugate; FITC, fluorescein isothiocyanate. GILGFVFTL-peptide pulsed cells were included as a positive control (J). Untreated cells were used as a negative control (A). Indicated is the percentage IFN--positive cells within the CD8<sup>+</sup>-T-cell population. The data are also presented as the number of IFN--positive spots, as measured in an IFN--specific ELISPOT assay (K). In addition, the percentage of specific lysis, as measured in chromium release assays, is shown (M). Effector cells were added at an effector/target cell ratio of 10, and specific lysis was calculated. A CTL clone specific for the NP<sub>418-426</sub> epitope was used as a control (L, N). The recognition of A/NL/95-M1 I59A and F64A could not be tested, since these mutant viruses could not be propagated to sufficiently high titers (*). Data from representative experiments are shown.
    </td></tr></tbody></table> </td></tr></tbody></table></center>
    <!-- null -->
    <center><table cellpadding="0" cellspacing="0" width="95%"><tbody><tr bgcolor="#e1e1e1"><td><table cellpadding="2" cellspacing="2"> <tbody><tr bgcolor="#e1e1e1"><td align="center" bgcolor="#ffffff" valign="top">
    View larger version (33K):
    <nobr>[in this window]
    [in a new window]
    </nobr> </td><td align="left" valign="top"> FIG. 5. Effect of selected amino acid substitutions in the M1<sub>58-66</sub> epitope on recognition by specific CTL. Reactivity of the CTL clone, directed against the M1<sub>58-66</sub> epitope, with stimulator cells infected with the wild type (WT) (B) or influenza virus A/NL/95-M1 I59L (C) or I59V (D) was determined by intracellular IFN- staining and flow cytometry. PE, phycoerythrin conjugate; FITC, fluorescein isothiocyanate. GILGFVFTL-peptide-pulsed cells were included as a positive control (E). Untreated cells were used as a negative control (A). Indicated is the percentage IFN-<sup>+</sup> cells within the CD8<sup>+</sup>-T-cell population. The data are also presented as the number of IFN--positive spots, as measured in an IFN--specific ELISPOT assay (F). In addition, the percentage of specific lysis, as measured in chromium release assays, is shown (H). Effector cells were added at different effector/target cell ratios as indicated, and specific lysis was calculated. A CTL clone specific for the NP<sub>418-426</sub> epitope was used as a control (G, I). Data from representative experiments are shown.
    </td></tr></tbody></table> </td></tr></tbody></table></center>
    Of four mutants tested, mutant virus A/NL/95-NP M426I was the<sup> </sup>only virus with an amino acid substitution in the NP<sub>418-426</sub><sup> </sup>epitope that proved viable and with which CTL recognition was<sup> </sup>studied. HLA-B*3501- and -A*0201-positive cells infected with<sup> </sup>wild-type virus were recognized by M1<sub>58-66</sub>-specific CTL and<sup> </sup>NP<sub>418-426</sub>-specific CTL by intracellular IFN- staining and ELISPOT<sup> </sup>and chromium release assays (Fig. 6B, E, and G). However, cells<sup> </sup>infected with A/NL/95-NP M426I were recognized by M1<sub>58-66</sub>-specific<sup> </sup>CTL, but not by NP<sub>418-426</sub>-specific CTL (Fig. 6C, F, and H).<sup> </sup>These results were confirmed by showing that the functional<sup> </sup>avidity of the NP<sub>418-426</sub>-specific CTL decreased more than 100-fold<sup> </sup>by the NP M426I substitution, using serial dilutions of wild-type<sup> </sup>and mutant peptide in ELISPOT assays (data not shown).<sup> </sup> <!-- null -->

    <center><table cellpadding="0" cellspacing="0" width="95%"><tbody><tr bgcolor="#e1e1e1"><td><table cellpadding="2" cellspacing="2"> <tbody><tr bgcolor="#e1e1e1"><td align="center" bgcolor="#ffffff" valign="top">
    View larger version (29K):
    <nobr>[in this window]
    [in a new window]
    </nobr> </td><td align="left" valign="top"> FIG. 6. Amino acid substitution NP M426I affects recognition by specific CTL. Reactivity of the CTL clone, directed against the NP<sub>418-426</sub> epitope, with stimulator cells infected with the wild type (B) or influenza virus A/NL/95-NP M426I (C) was determined by intracellular IFN- staining and flow cytometry. PE, phycoerythrin conjugate; FITC, fluorescein isothiocyanate. LPFEKSTVM-peptide-pulsed cells were included as a positive control (D). Untreated cells were used as a negative control (A). Indicated is the percentage of IFN--positive cells within the CD8<sup>+</sup>-T-cell population. The data are also presented as the number of IFN--positive spots, as measured in an IFN--specific ELISPOT (F). In addition, the percentage of specific lysis, as measured in chromium release assays, is shown (H). Effector cells were added at different effector/target cell ratios as indicated, and specific lysis was calculated. A CTL clone specific for the M1<sub>58-66</sub> epitope was used as a control (E, G). The recognition of A/NL/95-NP P419A, M426A, and P419G could not be tested, since these mutations prevented rescue of viable virus. Data from representative experiments are shown.
    </td></tr></tbody></table> </td></tr></tbody></table></center>
    The recognition of influenza virus A/NL/95-PB1 D593N was not<sup> </sup>tested, since this mutant virus could not be rescued. Although<sup> </sup>influenza virus A/NL/95-NP E46Q could not be propagated to sufficiently<sup> </sup>high titers, the recognition of HLA-A*0101-positive cells infected<sup> </sup>at a low MOI (0.02) was examined by ELISPOT. It was found that<sup> </sup>the substitution at the anchor residue abrogated recognition<sup> </sup>by NP<sub>44-52</sub>-specific CTL (data not shown). Influenza virus A/NL/95-NP<sup> </sup>R175K was not tested, since no specific CTL clone was available<sup> </sup>for this epitope.<sup> </sup> <!-- null -->
    <table bgcolor="#e1e1e1" cellpadding="0" cellspacing="0" width="100%"> <tbody><tr><td align="left" bgcolor="#ffffff" valign="middle" width="5%"></td> <th align="left" valign="middle" width="95%"> DISCUSSION </th></tr></tbody></table> <table align="right" border="1" cellpadding="5"><tbody><tr><th align="left"> Top
    Abstract
    Introduction
    Materials and Methods
    Results
    Discussion
    References
    </th></tr></tbody></table>
    In the present paper, the effect of amino acid substitutions<sup> </sup>in CTL epitopes on viral fitness and T-cell recognition was<sup> </sup>evaluated. It was concluded that functional constraints imposed<sup> </sup>on CTL epitopes limit escape from virus-specific CTL without<sup> </sup>loss of viral fitness.<sup> </sup>
    The synonymous/nonsynonymous analysis revealed that in the 90<sup> </sup>amino acids that constitute the 14 known epitopes located in<sup> </sup>the NP, relatively more nonsynonymous mutations occurred between<sup> </sup>1969 and 2003 than in the rest of the protein. The hypervariable<sup> </sup>epitope NP<sub>418-426</sub> had a major impact on the lower ds/dn ratio,<sup> </sup>and 5 out of the 14 partially overlapping epitopes were fully<sup> </sup>conserved. Some points in this analysis should be taken into<sup> </sup>consideration. First, since commonly old prototypic strains<sup> </sup>like A/Puerto Rico/8/34 have been used for the identification<sup> </sup>of influenza virus CTL epitopes, there is a bias towards the<sup> </sup>identification of conserved epitopes (12-14, 18, 25, 32, 33,<sup> </sup>54, 57, 59). Recent work in our laboratory indicates that a<sup> </sup>significant number of epitopes are not conserved (Berkhoff et<sup> </sup>al., unpublished data). Second, the conserved epitopes and the<sup> </sup>variable epitopes, including the NP<sub>418-426</sub> epitope, have in<sup> </sup>common that they all retained their anchor residues for binding<sup> </sup>to their corresponding HLA molecules. The only exception to<sup> </sup>this is an amino acid substitution at position 384 of the NP.<sup> </sup>The R384G substitution, which is at the anchor residues of the<sup> </sup>HLA-B*0801- and -B*2705-restricted epitopes NP<sub>380-388</sub> and NP<sub>383-391</sub>,<sup> </sup>resulted in the loss of their epitopes and abrogated recognition<sup> </sup>of virus-infected cells by specific CTL (51, 58). However, introduction<sup> </sup>of a glycine at position 384 of the NP of influenza virus A/Hong<sup> </sup>Kong/2/68 was detrimental to viral fitness, and several comutations<sup> </sup>associated with the R384G substitution in epidemic influenza<sup> </sup>virus strains were required to functionally compensate for the<sup> </sup>detrimental effect of the R384G substitution (50, 52). Similar<sup> </sup>findings have been observed for CTL escape mutants of HIV and<sup> </sup>simian immunodeficiency virus (SIV), which also accumulated<sup> </sup>extraepitopic comutations in the gag protein for restoration<sup> </sup>of viral fitness in the presence of mutations in CTL epitopes<sup> </sup>(17, 26, 42). Apparently, RNA viruses display sufficient flexibility<sup> </sup>to escape from CTL and retain viral fitness. For HIV and SIV,<sup> </sup>the selective pressure is mediated by CTL during the chronic<sup> </sup>infection of individual hosts, while for influenza viruses this<sup> </sup>takes place by CTL immunity at the population level (19). It<sup> </sup>is of special interest that also for HIV, CTL escape mutants<sup> </sup>can be identified at the population level (45), although transmission<sup> </sup>rates of this virus are much lower than those for influenza<sup> </sup>viruses. Thus, influenza virus CTL epitopes are either conserved,<sup> </sup>display variation at non-anchor residues, or lose their anchor<sup> </sup>residues at the cost of viral fitness, which is functionally<sup> </sup>compensated for by the accumulation of comutations. To assess<sup> </sup>the impact of amino acid substitutions in conserved epitopes<sup> </sup>on viral fitness and recognition by specific CTL, we conducted<sup> </sup>a mutational analysis of the epitope M1<sub>58-66</sub> (GILGFVFTL). This<sup> </sup>epitope is immunodominant and recognized by a large portion<sup> </sup>of individuals in the population, but is highly conserved. Replacement<sup> </sup>of the anchor residue at position 2 of the epitope (M1 I59A)<sup> </sup>was detrimental to viral fitness, whereas alanine replacements<sup> </sup>at the other eight positions did not prevent rescue of recombinant<sup> </sup>influenza virus and were tolerated to various extents. The M1<sub>58-66</sub><sup> </sup>epitope is located in the fourth N-terminal -helix of the M1<sup> </sup>protein. Mutations in this region may disturb the functional<sup> </sup>and structural integrity of the protein, as has been described<sup> </sup>for mutations in the M1 "helix six" domain (8, 31). The reduced<sup> </sup>virus titers obtained with a number of these mutant M1 viruses<sup> </sup>correlated with the number of productively infected cells, as<sup> </sup>measured by immunofluorescence assay using an NP-specific monoclonal<sup> </sup>antibody 6 h postinfection of MDCK cells, suggesting that the<sup> </sup>virus replication cycle was affected at an early pretranscriptional<sup> </sup>stage (data not shown). Conservative amino acid substitutions<sup> </sup>at position 2 of the M1<sub>58-66</sub> epitope (M1 I59L and I59V) were<sup> </sup>less critical, although the kinetics of viral replication was<sup> </sup>somewhat affected. More importantly, the A/NL/95-M1 I59L and<sup> </sup>I59V mutant viruses were fully recognized by M1<sub>58-66</sub> specific<sup> </sup>CTL, which makes it unlikely that these variants would ever<sup> </sup>emerge in the human population. Although some of the other alanine<sup> </sup>replacements resulted in the partial loss of recognition by<sup> </sup>M1<sub>58-66</sub>-specific CTL, their impaired replication kinetics is<sup> </sup>not in favor of the emergence of these mutants. We speculate<sup> </sup>that there must be a trade-off between viral fitness and immune<sup> </sup>recognition of which we have little insight at present. The<sup> </sup>T-cell recognition patterns that were observed here with mutant<sup> </sup>virus-infected cells were in agreement with those observed with<sup> </sup>mutant M1<sub>58-66</sub>-peptides in previous studies (1, 20, 41). Although<sup> </sup>the use of T-cell clones may not reflect the situation in vivo,<sup> </sup>the analysis of anchor residues boils down to recognition of<sup> </sup>the epitope or not, which is not different between clonal and<sup> </sup>polyclonal T-cell populations. In the analysis of T-cell receptor<sup> </sup>contact residues, as done for the M1<sub>58-66</sub> epitope, the situation<sup> </sup>is more complicated. However, the M1<sub>58-66</sub>-specific CTL response<sup> </sup>is oligoclonal in nature and dominated by T cells carrying the<sup> </sup>T-cell receptor with V? 17 chains (29, 37). Fitness<sup> </sup>costs also limit variation in the highly immunodominant Gag<sup> </sup>p11C, C-M CTL epitope of SIV and escape from specific CTL (43).<sup> </sup>Therefore, this phenomenon may be more universal and apply to<sup> </sup>more RNA viruses, which are under selective pressure mediated<sup> </sup>by CTL. It even may contribute to shaping of the T-cell repertoire<sup> </sup>and have an influence on the hierarchy of epitope dominance.<sup> </sup>
    Next, we wished to evaluate the conservative anchor residues<sup> </sup>of the otherwise hypervariable epitope NP<sub>418-426</sub> (LPFEKSTVM).<sup> </sup>The relatively conservative NP P419A and P419G substitutions<sup> </sup>at position 2 of the epitope were both detrimental to viral<sup> </sup>fitness, indicating that the proline at this position is essential.<sup> </sup>Amino acid substitutions at position 9 of the epitope, the second<sup> </sup>anchor residue, yielded interesting results. First, the NP M426A<sup> </sup>substitution was detrimental to viral fitness. Second, with<sup> </sup>the conservative NP M426I substitution, the HLA-B*3501 binding<sup> </sup>motif was retained (16, 23, 56) and viral fitness was not affected<sup> </sup>to a great extent. Of special interest, HLA-B*3501-positive<sup> </sup>cells infected with influenza virus A/NL/95-NP M426I were poorly<sup> </sup>recognized by NP<sub>418-426</sub>-specific T-cell clones. Since the NP<sup> </sup>M426I mutant epitope retained its capacity to bind to HLA-B*3501,<sup> </sup>it may have undergone conformational changes in T-cell receptor<sup> </sup>contact residues, preventing recognition by CTL, as has been<sup> </sup>described previously for another HLA-B*3501-restricted epitope<sup> </sup>(15). Conservative amino acid substitutions at the anchor residues<sup> </sup>of the epitopes PB1<sub>591-599</sub> (VSDGGPNLY), NP<sub>44-52</sub> (CTELKLSDY),<sup> </sup>and NP<sub>174-184</sub> (RRSGAAGAAVK) also affected viral fitness. The<sup> </sup>PB1 D593N substitution in particular was detrimental to viral<sup> </sup>fitness. Although the conservative NP E46Q substitution resulted<sup> </sup>in the loss of the anchor residue and would allow the virus<sup> </sup>to escape from specific CTL, the loss of viral fitness may limit<sup> </sup>the emergence of this variant in the human population.<sup> </sup>
    Based on the data presented here, we speculate that influenza<sup> </sup>A viruses display a limited degree of variation in CTL epitopes<sup> </sup>despite selective pressure on these epitopes mediated by CTL.<sup> </sup>Functional constraints imposed on influenza virus CTL epitopes<sup> </sup>may limit efficient escape from CTL and could constitute the<sup> </sup>Achilles heel of these viruses, limiting the impact of epidemic<sup> </sup>and pandemic outbreaks of influenza on severe morbidity and<sup> </sup>mortality.<sup> </sup>
    <sup> </sup>
    <!-- null -->
    Last edited by Anne; November 23, 2006, 04:25 PM.

    Comment


    • #3
      Re: Cell-mediated Protection in Influenza Infection

      free article

      Journal of Virology, May 2002, p. 4886-4890, Vol. 76, No. 10
      0022-538X/02/$04.00+0 DOI: 10.1128/JVI.76.10.4886-4890.2002
      Copyright ? 2002, American Society for Microbiology. All Rights Reserved.
      Protective Cross-Reactive Cellular Immunity to Lethal A/Goose/Guangdong/1/96-Like H5N1 Influenza Virus Is Correlated with the Proportion of Pulmonary CD8<sup>+</sup> T Cells Expressing Gamma Interferon

      Sang Heui Seo,<sup>1</sup> Malik Peiris,<sup>2</sup> and Robert G. Webster<sup>1</sup><sup>*</sup> Division of Virology, Department of Infectious Diseases, St. Jude Children's Research Hospital, Memphis, Tennessee 38105-2794,<sup>1</sup> Department of Microbiology, The University of Hong Kong, Queen Mary Hospital, Pokfulam, Hong Kong SAR, China<sup>2</sup>
      Received 2 November 2001/ Accepted 8 February 2002
      <!-- null -->
      <table bgcolor="#e1e1e1" cellpadding="0" cellspacing="0" width="100%"> <tbody><tr><td align="left" bgcolor="#ffffff" valign="middle" width="5%"></td> <th align="left" valign="middle" width="95%"> ABSTRACT </th></tr></tbody></table> <table align="right" border="1" cellpadding="5"><tbody><tr><th align="left"> Top
      Abstract
      Introduction
      Materials and Methods
      Results
      Discussion
      References
      </th></tr></tbody></table>
      A/Goose/Guangdong/1/96-like H5N1 influenza viruses now circulating<sup> </sup>in southeastern China differ genetically from the H5N1 viruses<sup> </sup>transmitted to humans in 1997 but were their precursors. Here<sup> </sup>we show that the currently circulating H9N2 influenza viruses<sup> </sup>provide chickens with cross-reactive protective immunity against<sup> </sup>the currently circulating H5N1 influenza viruses and that this<sup> </sup>protective immunity is closely related to the percentage of<sup> </sup>pulmonary CD8<sup>+</sup> T cells expressing gamma interferon (IFN-). In<sup> </sup>vivo depletion of T-cell subsets showed that the cross-reactive<sup> </sup>immunity was mediated by T cells bearing CD8<sup>+</sup> and T-cell receptor<sup> </sup>(TCR) /? and that the V?1 subset of TCR<sup> </sup>/? T cells had a dominant role in protective immunity.<sup> </sup>The protective immunity induced by infection with H9N2 virus<sup> </sup>declined with time, lasting as long as 100 days after immunization.<sup> </sup>Shedding of A/Goose/Guangdong/1/96-like H5N1 virus by immunized<sup> </sup>chickens also increased with the passage of time and thus may<sup> </sup>play a role in the perpetuation and spread of these highly pathogenic<sup> </sup>H5N1 influenza viruses. Our findings indicate that pulmonary<sup> </sup>cellular immunity may be very important in protecting na?ve<sup> </sup>natural hosts against lethal influenza viruses.<sup> </sup>
      <!-- null -->

      Comment


      • #4
        Re: Cell-mediated Protection in Influenza Infection

        In a preliminary experiment, we primed mice with the H1N1 PR8 strain and the H3N2 X31 strain followed by a challenge with A/Vietnam/1203/2004, one of the most lethal H5N1 viruses, which causes severe pathologic changes, even in ducks. While 9 of 10 naive mice died, 9 of 10 primed mice survived past day 10 of infectious challenge and recovered substantial weight
        I wonder how they "primed" the mice-via immunization? Mingus, might yuo know?

        I asume this is the study that prompted the recent comments that prior infection with other influenza strains may give some cross immunity to H5N1 (from my keyboard to God's ears!) or they are also reportign the same findings. One thing I asked about when that came up , and dont recall getting an answer on- is have the H5N1 vistims to date had prior flu strains? Do they have routine flu immunization available to them? Has anyone determined if they have been immunized at least once in the past. How long do antibodies and other markers remain detectable in serum?

        If that info becomes available , is there anyway to see if theres trends- like survivors are more likely to have antibodies to other strains in their system before H5N1 than non survivors (or vice versa or no correlation).

        I would bet that WHO or other labs have retained serum samples on many of the victims from samples they were sent for confirmation. Is it theroetically possible to do this on those samples?

        I know the sample size is likely to be a bit small for confidence, but whatever info we can find out may just help a lot during planning.
        Upon this gifted age, in its dark hour,
        Rains from the sky a meteoric shower
        Of facts....They lie unquestioned, uncombined.
        Wisdom enough to leech us of our ill
        Is daily spun, but there exists no loom
        To weave it into fabric..
        Edna St. Vincent Millay "Huntsman, What Quarry"
        All my posts to this forum are for fair use and educational purposes only.

        Comment


        • #5
          Re: Cell-mediated Protection in Influenza Infection

          I asume this is the study that prompted the recent comments that prior infection with other influenza strains may give some cross immunity to H5N1 (from my keyboard to God's ears!) or they are also reportign the same findings.
          This CTL mediated cross protective immunity has also been described in chicken. Infection with H9N2 Viruses possessing H5 like internal genes and circulating in the HK poultry markets protected agianst otherwise lethal HPAI H5N1 strains. Like in other reports, the immunization did not prevent virusshedding which underlines that the competent CD8 T lymphocytes target on infected cells exhibiting viral AGs on their surface and do not react directly with native virus.
          http://www.pubmedcentral.nih.gov/art...i?artid=115873

          The findings are consistent with the conclusions of ferguson in his mathematical approach suggesting a short-lived strain-transcending immunity http://cmbi.bjmu.edu.cn/news/report/2005/flu/106.pdf

          This hypothesis could also resolve the conundrum of this year's H5 spread among birds in Germany and neighboring areas in Europe. How could otherwise a highly lethal virus show a pattern of disseminated, small spottet outbreaks?

          In Austria there have been found humoral ab's in swans. These swans have probably infected with circulating LPAI strains prior to H5 infection providing sufficient CTL mediated resistance and time gain to develop Anti H5 Ab's.

          Comment


          • #6
            Re: Cell-mediated Protection in Influenza Infection

            <TABLE cellSpacing=0 cellPadding=0 width="100%"><TBODY><TR><TD>Eur J Immunol. 1996 Feb;26(2):335-9.</TD><TD align=right>Related Articles,<SCRIPT language=JavaScript1.2><!--var PopUpMenu2_LocalConfig_jsmenu3Config = [ ["ShowCloseIcon","yes"], ["Help","window.open('/entrez/query/static/popup.html','Links_Help','resizable=no,scrollbars= yes,toolbar=no,location=no,directories=no,status=n o,menubar=no,copyhistory=no,alwaysRaised=no,depend =no,width=400,height=500');"], ["TitleText"," Links "]]var jsmenu3Config = [ ["UseLocalConfig","jsmenu3Config","",""]]//--></SCRIPT><SCRIPT language=JavaScript1.2><!--var Menu8617300 = [ ["UseLocalConfig","jsmenu3Config","",""], ["Substance via MeSH","window.top.location='http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?itool=pubmed_Abstract&db=pubmed&cmd=Dis play&dopt=pubmed_pcsubstance_mesh&from_uid=8617300 '","",""], ["Cited in PMC","window.top.location='http://www.pubmedcentral.gov/tocrender.fcgi?action=cited&tool=pubmed&pubmedid=8 617300'","",""], ["Books","window.top.location='http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?itool=pubmed_Abstract&cmd=Retrieve&db=p ubmed&list_uids=8617300&dopt=Books'","",""], ["LinkOut","window.top.location='http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?itool=pubmed_Abstract&cmd=Retrieve&db=p ubmed&list_uids=8617300&dopt=ExternalLink'","",""]]//--></SCRIPT> Links </TD></TR></TBODY></TABLE><DD>
            An HLA-B35-restricted epitope modified at an anchor residue results in an antagonist peptide.

            Dong T, Boyd D, Rosenberg W, Alp N, Takiguchi M, McMichael A, Rowland-Jones S.

            Institute of Molecular Medicine, John Radcliffe Hospital, Oxford, GB.

            Peptides associated with HLA-B35 commonly have a proline or occasionally a serine residue in the P2 anchor position of the peptide, with a tyrosine at the C terminus. Based on this motif, we identified an octamer epitope from influenza A matrix protein which is presented by HLA-B35. The requirements for MHC binding and T cell receptor contact have been analyzed using analogs of this peptide with substitutions at positions 1, 2, 4, 7 and 8. The natural epitope contains a serine residue at P2 of the peptide. Substitution of this residue with proline (the favored amino acid in this position in B35-associated peptides) considerably enhances binding to HLA-B35 in the T2-B35 cell line, but the peptide is not recognized by the majority of CTL clones and can antagonize recognition of the index peptide. This suggests that a conservative substitution at the P2 anchor position results in a conformational change in the peptide-MHC surface exposed to the T cell receptor.
            </DD>

            Comment


            • #7
              Re: Cell-mediated Protection in Influenza Infection

              what's the advantage of this short lasting immunity ?
              Why does our body not just extend it ?
              I mean, why should human evolution establish an
              artificial time limit here which is not good for the host ?
              Nature was well capable of developing this other, long lasting immune-system. But why is this one too specific and can't recognize all flu ?
              I'm interested in expert panflu damage estimates
              my current links: http://bit.ly/hFI7H ILI-charts: http://bit.ly/CcRgT

              Comment


              • #8
                Re: Cell-mediated Protection in Influenza Infection

                Ganseerpel- thank you very much for your comments! I truly appreciate that, and the link to Fergusons study. That was quite well written, it answered questions Ive had about the potential semi-protective factor.

                Would it be fair to say that the LPAI strains may provide enough cross protection that the individual organism can then tolerate infection, and it appears the virus has changed to become less lethal- when its actually the population thats changed its susceptibility?

                I wonder how the new, widepread immunization in the endemic areas will affect the natural progression. The first thought that comes to mind is that it might artifically reduce the time it takes for a strain turnover. I wonder how that affects our chances of pandemic? I've been concerned widespread immunization allowed for more chances for other host species to be exposed to the virus, thus increasing the chances it would spawn a pandemic strain.

                But now I'm wondering that if the time to strain turnover is decreased by the avian population having a "herd CTL resistance", perhaps the inheriting strain may be less of a monster than H5N1. So that might be a good thing....if luck is with us?
                Upon this gifted age, in its dark hour,
                Rains from the sky a meteoric shower
                Of facts....They lie unquestioned, uncombined.
                Wisdom enough to leech us of our ill
                Is daily spun, but there exists no loom
                To weave it into fabric..
                Edna St. Vincent Millay "Huntsman, What Quarry"
                All my posts to this forum are for fair use and educational purposes only.

                Comment


                • #9
                  Re: Cell-mediated Protection in Influenza Infection

                  Infection with H9N2 Viruses possessing H5 like internal genes and circulating in the HK poultry markets protected agianst otherwise lethal HPAI H5N1 strains.
                  I read about this last winter, re: HK ducks - same situation with cross-reactive immunity from H9N2. I had wondered if humans had any flu VAX, within 2-3 weeks prior to H5N1 exposure, could that reduce mortality and severe respiratory and neurological complications?

                  I could envision this being done for front line HCW & essential infrastructure personnel.

                  .
                  "The next major advancement in the health of American people will be determined by what the individual is willing to do for himself"-- John Knowles, Former President of the Rockefeller Foundation

                  Comment


                  • #10
                    Re: Cell-mediated Protection in Influenza Infection

                    I guess, you need vaccine containing internal genes, not only HA and NA
                    I'm interested in expert panflu damage estimates
                    my current links: http://bit.ly/hFI7H ILI-charts: http://bit.ly/CcRgT

                    Comment


                    • #11
                      Re: Cell-mediated Protection in Influenza Infection

                      Would it be fair to say that the LPAI strains may provide enough cross protection that the individual organism can then tolerate infection,
                      One feature of the CTL mediated cross protective immunity is that it depends on highly variable genetic background.
                      The key factor of this system is its HLA dependance (?HLA restriction"), i.e. the (viral) antigen must be transported to the surface of infected cells and bind to a highly specific, individual-specific protein complex ("Major histocompatility complex" = MHC =HLA specific. class I molecule).
                      F.e. the immunodominant NP epitope 383-391 is ?HLA B27 restricted?, thus only the immune system of persons possessing this phenotype can recognise and handle this epitope. But only ca 5 % of the middle European population are HLA B27 positive, the rest of individuals will not show a reaction to this epitope.

                      The problem boils down to a complicate and nearly impredictible (on terms of present knowlegde) pattern of interaction limited by HLA- specific individual reactivity to numerous epitopes with a different HLA specific profile of different strains.

                      and it appears the virus has changed to become less lethal- when its actually the population thats changed its susceptibility?
                      A diminished lethality seems to be a evolutionary advantage in the progression of an epidemic.

                      I wonder how the new, widepread immunization in the endemic areas will affect the natural progression. The first thought that comes to mind is that it might artifically reduce the time it takes for a strain turnover. I wonder how that affects our chances of pandemic?
                      This cross prot. immunity has two major consequences: Host protection against otherwise potentially lethal strains, and ? perpetuation of the HPAI virus in a new host aquiring a new opportunity of evolution, recombination and reassortment -and accelerated emergence of new virulent H5N1 strains. Masking the disease signs of subsequent infections and prolonged virusshedding seem to be typical features in the course of infection.
                      what's the advantage of this short lasting immunity ?
                      Why does our body not just extend it ?
                      I mean, why should human evolution establish an
                      artificial time limit here which is not good for the host ?
                      Nature was well capable of developing this other, long lasting immune-system. But why is this one too specific and can't recognize all flu ?
                      The Mhc is estimated to be over 500 million years old and classical MHC molecules could be found
                      throughout the vertebrates. (Kaufman, 1994), (Bontrop, 1995), (Trowsdale, 1995), (Kasahara, 1995)
                      The great complexity of the human Mhc, gives the advantage of recognition of a wide variety of pathogens (for the price of potential overreactions)

                      I guess, you need vaccine containing internal genes, not only HA and NA
                      The antigen needs to be presented by an infected cell, therefore, with the common technique of inactivated vaccines it will probably not be possible

                      Comment


                      • #12
                        Re: Cell-mediated Protection in Influenza Infection

                        I don't feel that I well understand this all yet.

                        But do I have to understand ?
                        Can't we just ask our immunology expert for his pandemic
                        probability estimate ?
                        Does mankind have any immunulogical tricks to fight
                        a severe pandemic except those already discussed
                        vaccinations ?

                        I just need the probabilities. Only if I can't get them I
                        have to study the subject by myself.


                        LPAI : low pathogenic avian influenza
                        HPAI: high pathogenic avian influenza
                        MHC: major histocompatibility complex
                        CTL:cytotoxic T lymphocytes
                        HLA:human leukocyte antigen
                        F.e.

                        NP,HA,NA: flu genes
                        I'm interested in expert panflu damage estimates
                        my current links: http://bit.ly/hFI7H ILI-charts: http://bit.ly/CcRgT

                        Comment


                        • #13
                          Re: Cell-mediated Protection in Influenza Infection

                          what about the flu-mist ?
                          It uses whole viruses.
                          How are the chances that it would (temporarily ?)
                          give some protection against H5N1 ?
                          I'm interested in expert panflu damage estimates
                          my current links: http://bit.ly/hFI7H ILI-charts: http://bit.ly/CcRgT

                          Comment


                          • #14
                            Re: Cell-mediated Protection in Influenza Infection

                            In flumist the internal genes of a vaccine strain (A/Ann Arbor/6/60) are reassorted with the HA and NA from strains equvalent to the common inactivated vaccines.

                            It could help to protect against H5 infection depending on the homology of the internal genes of the challenge strain and the vaccine strain.

                            At least it protects from virusses within the same serotype. In a 2-year multicenter efficacy field trial it has been shown that the live vaccine protected also against an epidemic strain which was not well matched to the vaccine strains (H3/H3). http://www.ncbi.nlm.nih.gov/entrez/q...&dopt=Abstract

                            Comment


                            • #15
                              Re: Cell-mediated Protection in Influenza Infection

                              Immunity to H5N1 strains provided by CTL clones induced by prior human influenza A virus infection shown in adults living in US urban area

                              The Journal of Immunology, 1999, 162: 7578-7583.
                              Copyright ? 1999 by The American Association of Immunologists

                              Human CD8<SUP>+</SUP> and CD4<SUP>+</SUP> T Lymphocyte Memory to Influenza A Viruses of Swine and Avian Species

                              </NOBR><NOBR>Julie Jameson</NOBR>, <NOBR>John Cruz</NOBR>, <NOBR>Masanori Terajima</NOBR> and <NOBR>Francis A. Ennis<!-- null --><SUP>1</SUP></NOBR>
                              Center for Infectious Disease and Vaccine Research, University of Massachusetts Medical Center, Worcester, MA 01655
                              <!-- ABS -->Recently, an avian influenza A virus (A/Hong Kong/156/97, H5N1)<SUP> </SUP>was isolated from a young child who had a fatal influenza illness.<SUP> </SUP>All eight RNA segments were of avian origin. The H5 hemagglutinin<SUP> </SUP>is not recognized by neutralizing Abs present in humans as a<SUP> </SUP>result of infection with the human H1, H2, or H3 subtypes of<SUP> </SUP>influenza A viruses. Subsequently, five other deaths and several<SUP> </SUP>more human infections in Hong Kong were associated with this<SUP> </SUP>avian-derived virus. We investigated whether influenza A-specific<SUP> </SUP>human CD8<SUP>+</SUP> and CD4<SUP>+</SUP> T lymphocytes would recognize epitopes on<SUP> </SUP>influenza A virus strains derived from swine or avian species,<SUP> </SUP>including the 1997 H5N1 Hong Kong virus strains. Our results<SUP> </SUP>demonstrate that adults living in an urban area of the U.S.<SUP> </SUP>possess influenza A cross-serotype reactive CD8<SUP>+</SUP> and CD4<SUP>+</SUP> CTL<SUP> </SUP>that recognize multiple epitopes on influenza A viruses of other<SUP> </SUP>species. Bulk culture cytotoxicity was demonstrated against<SUP> </SUP>avian and human influenza A viruses. Enzyme-linked immunospot<SUP> </SUP>assays detected precursor CTL specific for both human CTL epitopes<SUP> </SUP>and the corresponding A/HK/97 viral sequences. We hypothesize<SUP> </SUP>that these cross-reactive CTL might provide partial protection<SUP> </SUP>to humans against novel influenza A virus strains introduced<SUP> </SUP>into humans from other species.<SUP> </SUP>
                              Full text:
                              http://www.jimmunol.org/cgi/content/...367c7afb4a4783

                              Comment

                              Working...
                              X