"Medicine, Faculty of"@en . "Medicine, Department of"@en . "Experimental Medicine, Division of"@en . "DSpace"@en . "UBCV"@en . "Kaan, Philomena Miewching"@en . "2009-09-23T19:34:37Z"@en . "2002"@en . "Doctor of Philosophy - PhD"@en . "University of British Columbia"@en . "Respiratory syncytial virus (RSV) is the most common cause of acute bronchiolitis in\r\ninfants. Alveolar macrophages (AM), cells which play major roles in lung defense mechanisms, are major targets for RSV infection in vivo and in vitro. This thesis examines the effects of cell maturation, age and sex of the host, and interaction of environmental particulates (PM10) on in\r\nvitro RSV infection of guinea pig AM. In addition, the expression of protein kinase in RSV-infected AM was studied. Electron microscopy localized immunogold labeled RSV antigens in the lysosomes of mature AM that restrict RSV replication while gold particles were observed in the cytoplasm of immature AM that support viral replication. Using RSV Yield as a measure of virus progeny per RSV immunopositive AM, immature AM from young animals showed a\r\nsignificantly higher RSV Yield compared to the same AM sub-population from adult animals. The data concerning animal gender showed that the distribution of AM sub-population from both sexes is similar but the RSV Yield of mature AM was greater from male than female guinea pigs. The introduction of PM10 during RSV infection of AM resulted in suppression of RSV Yield\r\nand RSV-induced cytokine production. The study of protein kinase expression identified Rsk, PKB and p70 S6K as potential candidates and MAPK and PKB as major pathways involved in RSV-mediated signal transduction in AM. In conclusion, the response of guinea pig AM to in vitro RSV infection is associated with expression of candidate protein kinases and is influenced\r\nby cell maturation, age and sex of the host animal, and interaction with extrinsic factors such as air pollution."@en . "https://circle.library.ubc.ca/rest/handle/2429/13081?expand=metadata"@en . "7802013 bytes"@en . "application/pdf"@en . "VITRO RESPIRATORY S Y N C Y T I A L VIRUS INFECTION IN GUINEA PIG A L V E O L A R MACROPHAGES by PHILOMENA MIEWCHING K A A N B. Sc., Simon Fraser University, 1985 M . Sc., University of British Columbia, 1995 A THESIS SUBMITTED IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY In THE F A C U L T Y OF G R A D U A T E STUDIES (Department of Experimental Medicine) Experimental Medicine Program We accept this thesis as conforming to the required standard THE UNIVERSITY OF BRITISH C O L U M B I A February 2002 \u00C2\u00A9 Philomena Miewching Kaan, 2002 In presenting this thesis in partial fulfilment of the requirements for an advanced degree at the University of British Columbia, I agree that the Library shall make it freely available for reference and study. I further agree that permission for extensive copying of this thesis for scholarly purposes may be granted by the head of my department or by his or her representatives. It is understood that copying or publication of this thesis for financial gain shall not be allowed without my written permission. Department of t X pe*i w1:100) provides further support for a protective role in these antibodies (50). In addition, trials of hyperimmune RSV globulin in high-risk infants have been considered successful in the prevention of RSV-induced lower respiratory tract infections (51,52). Cellular Immunity Due to obvious ethical reasons, there are limited data about the role of cellular immunity in RSV disease in humans. Evidence supporting a role of cell-mediated immunity in the pathogenesis of RSV bronchiolitis emerged from the unfortunate experience of the unsuccessful FI-RSV vaccine, and the study of RSV disease in immunocompromised individuals. Infants given the FI-RSV vaccine were not protected against subsequent natural RSV infection and many of them manifested unusually severe illness with high rates of hospitalization and mortality (53). Follow-up laboratory findings by Chin et al. (54) revealed lung infiltrates of neutrophils and eosinophils as well as blood eosinophilia, both of which were not associated with primary RSV infection and suggested an enhanced cell-mediated inflammatory (Th2 type) response (25). This hypothesis was supported by subsequent studies demonstrating a Th2 type response in FI-RSV vaccinated mice and a Thl type response in mice when challenged with wildtype RSV (55). However, alternative evidence suggests that cell-mediated immunity may play a central role in the clearance of established RSV infection. For example, cancer or bone marrow transplantation patients undergoing immunosuppressive chemotherapy have more severe RSV 9 infection and prolonged virus shedding (56-58). In addition, detection of RSV-specific cytotoxic T lymphocytes (CTL) in the peripheral blood of infected adults has been associated with decreased clinical symptoms (59, 60). Non-immunologic Factors Besides immunologic mechanisms, the anatomy of the lungs has been considered to play a role in severe RSV infection in infancy. By virtue of the relatively lower maximal expiratory flows at functional residual capacity, infants with smaller and narrower airways may be more prone to develop wheeze at the time of RSV bronchiolitis (40, 61). In addition, evidence from in vitro studies indicated that plasma proteins leaking into the airways as a result of inflammatory process may degrade and/or inhibit the function of pulmonary surfactant, which in turn could cause blockage of conducting airways and contribute to increased airway resistance (62, 63). Support for this possibility came from the work by Dargaville et al. (42) who demonstrated a deficiency of surfactant protein A with impaired functional activity in lavage fluids of infants with RSV bronchiolitis. Furthermore, administration of exogenous surfactant has been reported to alleviate disease symptoms in animal studies (37, 64) as well as in clinical studies (65, 66). 10 Important Issues At this point, I would like to draw attention to some issues in relation to RSV infections in humans. First, as discussed earlier, RSV infection during infancy can induce either a Thl type or Th2 type immune response depending on host and viral factors (67). However, classifying the immune response of RSV bronchiolitis in humans should be carried out with reservations. One should bear in mind that the dichotomy for activated T cells was originally described in the mouse (68)1 and the distinction between the 2 subsets of cells is not as definitive in humans. Secondly, the prevalence of asthma and atopy has been and is still on the increase over the past 3 decades (70, 71). This increase prevalence of respiratory allergies may be due to the interaction of independent factors (e.g., environmental and lifestyle) resulting in altered allergy-related immune responses (71). The Hygiene Hypothesis (72) may be a plausible explanation for the current unfavorable trend in asthma and atopy. The impact of the interaction between RSV and environmental particulates on A M function (see Chapter 8) may provide insights on this issue. 1 CD4+ T helper lymphocytes (ThO) play a major role in the modulation of immune responses. These ThO cells differentiate into 2 subclasses: Thl and Th2. Thl responses are usually observed in infectious diseases involving intracellular pathogens (e.g., bacteria and viruses) and delayed-typed hypersensitivity while Th2 responses deal with parasitic infections. Th2 responses are also associated with allergy-related diseases such as asthma. These subclasses of cells are distinguished by their different cytokine profiles: Thl cells produce IL-2 and IFNy, both of which mediate help for the generation of cytotoxic T cells. Th2 cells secrete IL-4, IL-5, IL-10 and IL-13, cytokines which induce B cells to produce specific immunoglobulins (69). 11 ANIMAL MODELS OF R S V INFECTION Although it is not clear that the immune response demonstrated in animals can be directly extrapolated to human infection, animal studies have contributed much to the understanding of immunopathogenic mechanisms in RSV disease. These in vivo systems have facilitated the testing of many hypotheses that would have been otherwise deemed unethical or impractical i f experiments were attempted in humans. Since the isolation of RSV, numerous animal models of RSV infection have been developed using primates, cotton rats, mice, calves, sheep, guinea pigs, rats, ferrets and hamsters (73-75). As in humans, the manifestation of RSV disease in different animal models varies with age, genetic makeup and immunologic status (73). In experiments using primates, only the chimpanzee appears to be highly permissive to RSV replication even though other primate species could be infected by intranasal instillation of RSV (13). Among the animal species used as models of lower respiratory tract infection with RSV, one popular animal model is the Balb/c mouse. While RSV infection produces asymptomatic bronchiolitis, this model has been useful in the study of cell-mediated immunity in RSV infections (76). This model system provided the evidence that the cell-mediated enhancement of RSV disease seen in humans was due to an imbalance of Thl and Th2 lymphocyte response to the FI-RSV vaccine. In the murine model, RSV induces a Thl type CD4+ response characterized by production of Thl cytokines (IFNy, IL-2, IL-12) and cytotoxic T cells (55, 59, 77). In contrast, immunization with inactivated virus or G subunit viral proteins induces a Th2 CD4+ response with the corresponding cytokines (IL-4 and IL-5) and no C T L (55). In addition, studies with mice also indicated that RSV-specific CD8+ lymphocytes can clear virus from persistently infected, immunodeficient mice, but acute pulmonary damage occurs and may be associated with fatality (67, 78). Thus, both CD4+ and CD8+ lymphocytes are involved in both recovery and the pathologic response even though CD4+ cells have been demonstrated to be more antiviral (i.e., 4-fold less in cell number to 12 decrease lung virus titer and produce similar effects) and more immunopathogenic than CD8+ cells in RSV infected mice (79). The guinea pig model for RSV infection is the basis of RSV research in our laboratory and in this thesis. Besides the extensive knowledge of airway physiology in this species, Hegele et al. (80, 81) were the first to successfully induce acute and persistent R S V infection and demonstrated RSV genome as well as viral antigens in airway epithelial cells and A M from RSV-infected guinea pigs. Riedel and coworkers (82) demonstrated hyperresponsiveness to histamine challenge and viral persistence in RSV-infected guinea pigs even after resolution of virus-induced inflammation in the lung. Moreover, Dakhama et al. (83) demonstrated that R S V could infect and replicate in guinea pig A M in cell culture medium. RSV INFECTION IN VITRO In vitro experiments permit the isolation and manipulation of a microenvironment in a controlled and systematic manner. In most diagnostic virology laboratories, RSV infection is determined by rapid antigen detection test kits (for example, enzyme immunoabsorbent assay (EIA) kits, and immunofluorescent techniques). However, the detection of infectious virus by cell culture remains the gold standard. This technique is used throughout this thesis for detection of replicating RSV as well as quantification of viral titers. RSV-sensitive, transformed human epithelial cells, HEp-2 cells were used for RSV detection because of their characteristic cytopathic effect of syncytia formation as a result of the fusion of cell membranes between infected and uninfected adjacent cells. Epithelial Cells Since the airway epithelial cell is the primary target for RSV infection, it is speculated that these cells and the soluble factors (cytokines and mediators) derived from them are 13 responsible for initiation of airway immune responses and inflammatory processes (84, 85). In epithelial cells, viral replication causes upregulation of multiple cytokines, chemokines and adhesion molecules by triggering intracellular signaling pathways (86-88). In vitro infection in epithelial cells with RSV activates transcription factors such as activator protein-1 (AP-1), nuclear factor-kappa B (NF-KB ) , NF-IL6 and CCAAT-enhancer binding proteins (c/EBP) (89, 90). A major cytokine induced by RSV infection is interleukin-8 (IL-8) (91, 92). IL-8 is a potent chemotactic agent for neutrophils (93). Clinical studies have demonstrated that lavage fluid from RSV-infected infants contains a predominance of neutrophils (94). Another cytokine that is worthy of mention is interleukin-6 (IL-6). IL-6 is detected in respiratory secretions during RSV infection in vivo (95) as well as in RSV infected epithelial cell cultures in vitro (95-98). During RSV infection IL-6 may play an immunoregulatory role by promoting humoral and cellular defense mechanisms (99). IL-6 plays a major role involved in isotype switch, differentiation and synthesis of IgA (100). On the other hand, IL-6 may play a proinflammatory role by contributing to the symptoms and signs of acute infections (99). Therefore it appears that RSV may cause inflammation in part through the production of these proinflammatory cytokines from epithelial cells in the lung. In addition to the accumulation of neutrophils, eosinophils have been demonstrated in the inflammatory infiltrates from autopsied lungs of children infected with RSV (53). Studies by Becker et al. (101) demonstrated that RSV infection of human airway epithelial cells caused the production of an important eosinophil chemoattractant, RANTES (Regulated upon activation, normal T cells expressed and presumably secreted). Garofalo et al. (47) demonstrated the increased levels of ECP in lavage fluid sampled from RSV-infected infants. Moreover, subsequent studies have indicated that degranulation of the accumulated eosinophils and the release of ECP may be mediated via upregulation of C D 18 molecules on the R S V infected epithelial cells (102). 14 Other cytokines that have been implicated in the promotion of the inflammatory response induced by RSV infection include interferon-p4 (IFN(3), interleukin-la ( IL-la), macrophage inflammatory protein-la (MIP-loc) and monocyte chemotactic protein (MCP) (92, 102, 103). Enhanced expression of IFNP and I L - l a causes upregulation of the major histocompatible complex (MHC) class I molecules on the surface of the epithelium, which in turn results in activation of CD8+ T cells and increases local inflammation (103, 104). M l P - l a and M C P are known chemotactic and activator factors for monocytes, basophils and eosinophils. In addition, M l P - l a has been shown to stimulate the production of IgE and IgG4 by human B cells (105) and to induce degranulation of natural killer cells (106), thereby contributing to the augmentation of local inflammation. Alveolar Macrophages Besides epithelial cells, A M (107, 108), eosinophils (109), T cells and neutrophils (110) are also cell targets for RSV infection. A M are the first line of defense against inhaled pathogens (biologic and non-biologic) and are in direct encounter with the virus during RSV infection. In addition, A M are the pivotal cell type that link the external environment to within the lung. For these reasons, A M are very important and are the cell type of focus in this thesis work. Panuska et al. examined bronchoalveolar lavage cells from adult transplant patients with RSV infection and found both A M and epithelial cells were infected with RSV in vivo (107). Studies investigating susceptibility of blood monocytes, A M and cord blood to RSV infection indicated a role for intrinsic cellular factors in the restriction of viral infection (83, 108). Panuska et al. ( I l l ) demonstrated that human A M are able to support prolonged RSV replication in vitro. Although A M are susceptible to R S V infection, in comparison to epithelial cells, the infection is usually abortive after the initial cycles of viral replication (112,113). In addition, RSV does not 15 induce cytopathic effects in A M and only a fraction or subpopulation of A M was susceptible to RSV regardless of the dose of infectious virus used. This observation has also been reported in mice (114), cattle (115) as well as in guinea pig (83). The effects of in vitro RSV infection on A M have been studied (83, 111, 114). Exposure of A M to RSV has been reported to stimulate the immunoregulatory functions and depress the microbicidal activity of these cells (114). With regards to cytokine expression, as demonstrated in epithelial cells, RSV induced upregulation of IL-1 (114), IL-6 (114, 116, 117) and IL-8 (116, 117) in A M . Of particular interest is tumor necrosis factor-oc (TNFa), a marker of macrophage activation. T N F a is induced in RSV-infected A M but not expressed in epithelial cells infected by RSV (91, 114, 116, 117). T N F a is synthesized and secreted in large quantities by macrophages in response to various proinflammatory stimuli (e.g., lipopolysaccharide). More importantly, T N F a is mediated by N F - K B activation and induces macrophage synthesis of several key proinflammatory cytokines (e.g., IL-1, IL-6, GM-CSF), chemokines (e.g., RANTES, IL-8, MCP-1 and M l P - l a ) as well as inflammatory mediators (e.g., leukotrienes, reactive oxygen species). The combined effects of these mediators on cell recruitment and cell activation are likely to play a fundamental role in the pathogenesis of RSV disease. In contrast to epithelial cells, there is limited published literature on how RSV modulates these mediators via intracellular signaling pathways in A M (see Chapter 9). SUMMARY This year, 2002, marks the 46 t h anniversary since the discovery of RSV. While the impact of RSV from public health and economic perspectives is appreciable, there are numerous issues that remain controversial. These include the pathogenic/ protective role of RSV against development of allergies in later life, the role of RSV in reactive airway disease, and whether 16 RSV infection induces a Thl or Th2 response in humans. Epidemiological studies suggest that the influence of viral infection on the developing immune system depends on the type of viral infection, the age, sex, and genetic background of the host and environmental factors. The purpose of this thesis is to examine the relationship between RSV infection and anti-viral immune response by evaluating some of these factors in a controlled environment in order to better understand the nature of RSV disease. 17 C H A P T E R 3 A L V E O L A R M A C R O P H A G E S INTRODUCTION The alveolar macrophages (AM) play major roles in the lungs defense against bacterial infections (e.g., Mycobacterium, Pseudomonas), parasitic infections (e.g., Leishmania), viral infections (e.g., HIV, RSV), chronic inflammation (e.g. particulates, silicosis, asbestosis) and in response to cigarette smoke and air pollutants (118). As \"gatekeepers\", A M are responsible for clearance of microorganisms and small inhaled particulate material in the alveolar space. The aim of this chapter is to provide a basic review of A M biology in general and highlight issues that are relevant to RSV infection in A M . ORIGIN OF MACROPHAGES Macrophages originate from a pluripotent stem cell (colony-forming unit granulocyte-macrophage (CFU-GM)) in the bone marrow (119, 120). In response to growth factors, such as interleukin-3 (IL-3), granulocyte-macrophage colony stimulating factor (GM-CSF) and macrophage colony stimulating factor (M-CSF), the pluripotent stem cells from the bone marrow develop into myeloid stem cells that eventually become promonocytes (121, 122). As the promonocytes mature, they leave the bone marrow and enter the blood stream to become blood monocytes. Monocytes circulate in the blood stream for 1 - 2 days and continue to mature within the blood stream. Monocytes are randomly distributed and recruited into different organs, i.e. Kupffer cells in the liver, Langerhan cells in the skin, osteoclast cells in bone, microglial cells in the central nervous system, peritoneal macrophages in the intestine and pulmonary macrophages in the lung. In lung capillaries, blood monocytes traverse the blood vessel wall by a process known as diapedesis and mature into pulmonary macrophages (123). Recruitment 18 from the bone marrow is the dominant mechanism for renewal of the pulmonary macrophage population. This recruitment process is mediated by monocyte-specific chemoattractants including MIP-1, MIP-2 and monocyte chemoattractant protein-1 (MCP-1) (124, 125). In addition, a very low percentage (0.5%)_pf A M proliferates locally and contributes to the alveolar pool of cells (126). In vitro experiments have demonstrated that human A M are capable of proliferation in response to GM-CSF (127). HETEROGENEITY OF MACROPHAGES Macrophages are tissue-specific and in different tissues they are morphologically and functionally distinct depending on the microenvironment in which they mature. However, the basic functions of macrophages remain similar despite their ability to excel in specific functions at various body sites. For example, A M are excellent at phagocytosis and secretion of nitric oxide (NO) but are somewhat deficient at antigen presentation. In contrast, macrophages in the spleen and lymph nodes are excellent at antigen presentation. Macrophage heterogeneity is therefore a reflection of the local microenvironment of the cell and its involvement in various physiological or pathological processes (118). Pulmonary macrophages are the only macrophages living in an aerobic environment and are categorized into 4 different types based on their localization: intravascular macrophage, dendritic cell, interstitial macrophage and A M (128). Intravascular macrophages are located on the endothelial cells of the pulmonary capillaries (129). These cells are highly phagocytic and are responsible for clearance of pathogens entering the lung via the blood stream (130). Apparently, these cells have been described in humans, pigs, cats, dogs and sheep but not in rodents (131). Dendritic cells, located in the lung interstitial tissue, are poor phagocytes (132, 133). High amounts of Class II antigen expression on the surface of dendritic cells make them suited for antigen presentation and accessory function (128). Interstitial macrophages are located 19 in the lung connective tissue. Like A M , interstitial macrophages are effective phagocytes, function as antigen presenting cells as well as produce cytokines (e.g., TNFa, IFNa/P) and oxygen radicals (131). Unlike A M , they have an increased ability to replicate and synthesize D N A in vitro (134, 135) and are much more effective in stimulating T-cell responses to antigens (136). Morphometric comparisons of A M , interstitial macrophages and blood monocytes indicated that interstitial macrophages represent a transitional stage of maturation between blood monocytes and A M (137). A M are localized on the alveolar epithelial surface within a film of surfactant. As a first line of defense, A M possess high phagocytic and microbicidal potential. In addition, A M are also involved in inflammatory activities and regulation of lung homeostasis (see below) (136). Since A M can be easily obtained by bronchoalveolar lavage (>95% purity), in comparison to other types of pulmonary macrophages, these cells have been most extensively studied (138). A M that are newly recruited into the alveolar space continue to undergo maturation (136). A M represent a morphologically and functionally heterogeneous population. As A M mature, they increase in size; ranging from 12 | im (generally the newly recruited cells) to 40 | im (the older and sometimes multinucleated A M ) in diameter (139). Taking advantage of these size and density differences, several investigators (140-142) have fractionated A M into different subpopulations using density gradient centrifugation and characterized these subpopulations and studied their heterogeneity based on receptor expression, release of mediators such as oxygen radicals and tissue factor, cytotoxicity, migration and pinocytosis (140,143, 144). Likewise, the methodology of isolating A M into subpopulations has been applied in 3 studies (Chapters 5, 6 and 7) in this thesis work. In addition, Sibille et al. (131) have established that the response of each subpopulation of A M to different stimuli (including response to RSV - see Chapters 5 and 6) may be markedly heterogeneous. In humans, the distribution of A M subpopulations is known to vary with different disease states. For example, the predominance of small monocyte-like A M 20 has been observed in acute inflammation whereas an increase in larger and more mature A M is associated with chronic lung disorders (145). MORPHOLOGY OF A M Light Microscopy A M are large irregular shaped cells with prominent ruffled membranes ranging from 12-40 um in diameter. The kidney-shaped nucleus, usually eccentrically placed within the cell, contains fine nuclear chromatin with one or two prominent nucleoli. Newly recruited A M have a relatively smaller cytoplasm-to-nucleus ratio while mature A M have a larger ratio due to an abundant cytoplasm with an increased number of cytoplasmic inclusions (146). Besides cytoplasmic vacuoles, fine granules and multiple large azurophilic granules are seen in the cytoplasm (147). Electron Microscopy Electron microscopy shows the multi-lobulated nucleus with fine clumps of chromatin. The cytoplasm contains well-developed Golgi complex, numerous vesicles, vacuoles and pinocytic vesicles, large mitochondria and lysosomes. Membrane bound, electron dense vesicles filled with ingested biological or inert particulate material are also visible. The ruffled cell membrane of the A M exhibits irregular microvilli or pseudopods that are involved in amoeboid movement and phagocytosis (147). FUNCTIONS OF A M A M perform their function as \"gatekeepers\" mainly through phagocytosis (when the alveolar space is invaded by inhaled pathogens), recruitment of inflammatory cells (in response to tissue and vascular changes), and regulation of lung homeostasis. These functions are carried 21 out and assisted by the wide range of surface receptors and large spectrum of secretory products within membrane bound organelles within the A M . Phagocytosis As the resident phagocyte of the alveolar space, A M play major role in defending the host against invasion by a wide variety of small particles (0.5 | im to 3.0 |j.m diameter) that are inhaled into the distal airways (148). These inhaled substances may be of biological (for example, bacteria, viruses, fungi and protozoa) as well as non-biological (for example, asbestos, particulate material) nature (128). Since A M are among the first cells encountering the virus in the event of a RSV infection, these cells play a major role as a first line of defense. However, and more importantly, A M trigger the immune response cascade through activation of granulocytes (see section on Inflammatory Activity). In addition to invading pathogens, A M also play an important part in scavenging apoptotic, senescent and damaged cells. Through phagocytosis and the production of oxygen radicals and proteases, most of the microorganisms and particulates are eliminated by the A M (128). Phagocytosis is a special form of endocytosis; unlike pinocytosis, a constitutive process that occurs continuously, phagocytosis is a receptor-mediated process that requires the transmission of extracellular signals to the interior of the cell to initiate the response (149). The main features of phagocytosis are attachment, engulfment, phagosome formation and maturation, digestion and membrane retrieval (118). A M move toward inhaled particles guided by a gradient of chemotactic molecules. Particle attachment and engulfment may be facilitated by the presence of specific receptors on the surface of the cell and opsonins that coat the inhaled particles. Several classes of receptors that promote phagocytosis have been characterized (see Table 3.1). 22 Table 3.1 Ligands recognized by A M via receptors (Extracted from Lohman-Matthes M L et al. (128), with permission) Immunoglobulins IgGl , IgG2a (murine) IgG2b, IgG3 (murine) IgGl , IgG3 monomers (human) IgG complexes (human) IgE, IgA (murine and human) Complement C3b, iC3b, C4b, C3d, C5a Lipoproteins Low density lipoproteins P-very low density lipoproteins Proteins Fibronectin, fibrin Lactoferrin, transferrin GM-CSF, CSF-1 Interferon-y, IL-4, IL-IRa IL-2, insulin Lectins with specificity for oc-linked galactose residues N-acetylgalactosamine residues N-acetylglucosamine residues a-linked fucose residues Mannose residues N-acetylneuramine residues Surface markers Class II molecules, C D l l a , C D l l b , C D l l c , CD14, CD18, CD54 Molecules recognized by monoclonal antibodies: 25F9 (mature macrophages), 27E10 (inflammatory macrophages), K i M2, K i M8 (mature macrophages), RM31 (inflammatory macrophages), RFD1, RFD7, RFD9 Ig: immunoglobulin; GM-CSF: granulocyte-macrophage colony stimulating factor; IL: interleukin Three major groups of receptors that play important roles in the attachment stage of phagocytosis are: the Fc receptors (which recognize the Fc region of antibodies coating inhaled pathogens), the complement receptors (which recognize the cleaved products of the third component of the complement C3) and lectin-binding receptors (which recognize lectins on microorganisms) (128). However, the specific A M receptor(s) that recognize RSV is yet to be determined, although the toll-like receptors are thought to be involved (150). Opsonization is an essential mechanism that enhances the initial stages of phagocytosis. Opsonins were initially discovered as factors that \"butter\" the particles to make them \"appetizing\" to phagocytes (151). Opsonins are soluble components in the alveolar space that facilitate phagocytosis by increasing the recognition and endocytosis of different pathogens by the A M through their binding to a specific receptor on the cell surface (131). Examples of opsonins include surfactant phospholipids, glycoproteins (e.g., fibronectin), proteins (e.g., IgG, IgA) and complement fragments (C3b) 23 (131). For example, antibodies may coat inhaled pathogens so that the Fc regions are exposed on the exterior. When the Fc portions of antibodies attach to the Fc receptors on the surface of A M , intracellular signals are triggered. These signals initiate the reorganization of cytoskeletal microfilaments at the site of attachment and result in the extension and fusion of the A M pseudopods and consequently engulfment of the inhaled particle (149). A number of hypotheses have been proposed for the mechanisms involved in the engulfment stage of phagocytosis. In Fc- and complement receptor-mediated uptake of inhaled particles, a \"zipper\" mechanism guided macrophage pseudopods to circumferentially envelop the attached particle (152). The bacterium Legionella pneumophilia enters the macrophage via an alternative coiling mechanism in which the microbial outer wall components induce the cell membrane to wrap around the organism in myelin-like configurations (153). Alternatively, engulfment could also occur independent of opsonins; for example, certain bacteria possess surface determinants or carbohydrate residues that interact with the mannose-fucose receptor on the surface of the A M (154). Engulfed particles are contained in large membrane-bound endocytic vesicles called phagosomes within the cytoplasm. The diameters of the phagosomes are determined by the size of the ingested particle and could be as large as the A M itself (149). At this stage, recycling of membrane and receptors takes place via sorting endosomes which are acidic vesicles that dissociate receptors from ligands and are responsible for shuttling of cell membrane and receptors back to the cell surface (155). The phagosomes containing the ingested particles fuse with primary lysosomes and mature into secondary lysosomes or phagolysosomes (156). The primary lysosomes contain lysosomal enzymes, which comprise more than 40 different acid hydrolases and are optimally active at a pH of about 5, that are involved in the degradation of phagocytosed material (131). In addition, A M possess granules containing oxygen metabolites (such as superoxide anion (CV), hydrogen peroxide (H2O2) and hydroxyl radical (OH')) and other enzymes that can be released into the phagosomes or the external 24 environment during phagocytosis to destroy ingested materials (see Table 3.2) (157). A M are equipped with various ways to degrade and eliminate ingested matter. Degradation of ingested matter in the A M depends on the nature of the ingested substance and/or the opsonins that are involved. For example, Fc-mediated phagocytosis results in the release of oxygen radicals and arachidonic acid metabolites (158, 159). In contrast, complement receptor-mediated phagocytosis does not involve oxygen or arachidonic metabolites but induces secretion of IL-1 (160). Table 3.2 Major products released by A M (Extracted and modified from Sibille Y et al. (131), with permission) Proteins Enzymes Antiproteases (e.g.al-proteinase inhibitor) Lysozyme, (^-glucuronidase Other inhibitors (e.g. IL-1 inhibitor) Acid hydrolase Glycoprotein (e.g. fibronectin) Angiotensin converting enzyme Complement components (e.g. C2, C4) Elastase: serine and metalloenzyme Binding protein (e.g. transferrin, ferritin) Collagenase (fibroblast-like, gelatinase) Free fatty acids Plasminogen activator Antioxidants (e.g. Glutathione) Cysteine proteinase (cathepsin L) Coagulation factors (e.g. Factors V , VII) Cytokines Biologically active lipids Interleukin-1 a, Interleukin-1 P Thromboxane A2 Interleukin-6 Prostaglandins Tumor necrosis factor 5-hydroxyeicosatetraenoic acid Interferon-cc, Interferon-y Leukotrienes Colony stimulating growth factors Platelet activating factor Transformation growth factor-P Fibroblast growth factor, insulin growth factor-1 Neutrophil-activating factor Enzyme-releasing peptide Neutrophil chemotactic factor Platelet-derived growth factor Histamine releasing factor 25 Following degradation of the ingested material, A M are removed from the alveoli by a number of routes. Some A M are transported up the airway by the mucociliary escalator while other A M migrate to the regional lymph nodes via peribronchial lymphatics. In addition, pathogen-containing A M can also be scavenged and phagocytosed by other A M (131). Although A M are well-equipped with a large spectrum of enzymes with microbicidal activities for destruction of many microorganisms, certain pathogens have evolved anti-immune strategies to counteract macrophage defenses in order to survive, parasitize and replicate within these cells. For example, Mycobacterium tuberculosis releases sulpholipids that prevent the fusion of lysosomes and phagosomes thereby avoiding exposure to the lysosomal enzymes (161). The Epstein-Barr virus (EBV) synthesizes a homologue of interleukin-10 (IL-10) so that its anti-inflammatory properties are retained; however the ability to stimulate lymphocytes is lost (162). While it has been postulated that RSV induces IL-10 production in A M to prevent complete anti-viral immunity (163); direct evidence is lacking and further investigations are needed to provide insights on the evasion strategies used by this virus to circumvent macrophage defenses. Regardless of the evasion strategy utilized by RSV, this virus is capable of causing productive infection in isolated human A M (111). Inflammatory Activity A M play an important role in the recruitment of additional inflammatory cells during the initial phase of lung inflammatory response when the load of inhaled material is overwhelmed and resident A M are insufficient to mount an adequate response (136). In the lung of the normal individual, A M are a major source of chemotactic factors (131). A M release various chemotactic factors to recruit inflammatory cells such as polymorphonuclear leukocytes (PMN), monocytes and lymphocytes. In the event of a RSV infection, A M play major roles in the activation and recruitment of granulocytes with P M N being the first cells recruited to the site of infection (164). 26 On interaction with pathogens or other stimuli such as complement, A M become stimulated and release chemotactic factors that attract P M N from the capillary lumen into the alveolar space. There are numerous macrophage-derived chemotactic factors for P M N (131); however, a few that are of particular interest include leukotriene B4 (LTB4) (165), TNFoc (166), interleukin-8 (IL-8) (167,168), and MIP-la(169). Upon stimulation by IgG or IgE immune complexes, A M release several arachidonic acid metabolites including LTB4. As one of the earliest factors identified, LTB4 is considered as the predominant P M N chemotactic factor secreted by human A M (165). LTB4 is a potent proinflammatory mediator and chemoattractant for P M N as well as eosinophils (170, 171). LTB4 increases pulmonary vessel permeability by inducing plasma leakage and leukocyte adhesion (172). In addition, L T B 4 in combination with other mediators (e.g., PAF and 5-HETE), mediates enzyme degranulation of P M N (173). During phagocytosis or stimulation by calcium ionophore (174), LPS and other microbial agents, IL-2, GM-CSF and IL-1 stimulate A M to produce TNFoc, a potent chemotactic factor for monocytes and P M N (166). TNFoc was initially identified for its ability to induce hemorrhagic necrosis of certain tumors (175). Subsequently, TNFoc was shown to have cytotoxic effects in vitro against various human tumors. TNFoc is involved in the induction of IL-1, GM-CSF, IL-6 and intercellular adhesion molecule-1 (ICAM-1) production (176). In addition, TNFoc activates P M N , enhances their phagocytic capabilities, and induces secretion of H2O2 (177, 178). Moreover, TNFoc is responsible for monocyte differentiation and the subsequent release of IL-1 from these cells (179). IL-1 stimulates the release of granule proteins (lysozyme and lactoferrin) in P M N (178). TNFoc induces upregulation of IL-6, which in turn induces the proliferation of immature and mature T cells (180). IL-6 also acts on B-lymphocytes as a differentiation factor and induces immunoglobulin secretion (181,182). 27 LPS stimulation of A M causes the release of IL-8 (168), a potent chemoattractant for P M N and T cells (183, 184). In addition to recruitment of P M N to sites of infection, IL-8 also regulates P M N activation. IL-8 increases the microbicidal ability of P M N by enhancing phagocytosis, O2\" generation and granule release (185). Indeed, production of these proinflammatory cytokines (i.e. TNFa, IL-8 and IL-6) has been enhanced during RSV infection (116). M l P - l a is produced by A M as well as P M N and epithelial cells during lung inflammation (186). M l P - l a has been shown to have chemoattractant activity for T cells, eosinophils and basophils (187). M l P - l a also plays a role in regulation of hematopoiesis through stimulation of other inflammatory mediators (e.g., TNFa, IL-1, IL-6, histamine). Increased M l P - l a expression has been observed in animal models of bacterial sepsis, silicosis and oxidant-induced lung injury (188-190). In addition, recent studies have demonstrated high levels of M l P - l a expression in A M from asthmatic individuals (191, 192). Recent studies have demonstrated that in vitro RSV infection of A M stimulated the expression of M l P - l a (193). Moreover, increased expression of M l P - l a has recently been associated with severe RSV bronchiolitis in infants (194). Regulation of lung homeostasis During inflammation, A M may elaborate many proinflammatory factors in concentrations that result in irreversible injury to the lung. In addition, A M may release cytokines and other factors that affect the function of other cells such as epithelial cells and fibroblasts (136). A M play an important role in the maintenance of lung homeostasis by the secretion of various inhibitory factors against inflammatory cells. For example, A M can inhibit P M N activity through the release of prostaglandin E2 (PGE2). In vitro studies have shown that 28 PGE2 can inhibit P M N and monocyte chemotaxis as well as the release of O2\" anions by P M N (195). Furthermore, recent studies by Thomassen et al. (196) have demonstrated that A M may regulate lung homeostasis via the release of N O which downregulate A M production of proinflammatory cytokines T N F a and M l P - l a . SUMMARY A M play a pivotal role in the innate immune response in defending the host against invading pathogens. As a first line of defense, A M limit the spread of infection and stimulate the immune response cascade through activation of granulocytes as well as activation of cytotoxic and helper T cells and the production of specific immunoglobulins. Despite the tightly regulated mechanisms against invaders, the RSV has devised strategies to not only evade subsequent immunosurveillance but also to survive within infected A M with productive replication of virus progeny. The study of RSV infection in different A M subpopulations and the viral impact on cytokine production as well as activation of signaling cascades in these cells will further our understanding of A M functions. 29 CHAPTER 4 GENERAL MATERIALS & METHODS C E L L CULTURE A l l cell culture work was carried out in a Biosafety cabinet (Forma Scientific, Marietta, OH) housed in a Level II containment facility. A l l solutions and equipment that come into contact with cells were steam sterilized or purchased as \"sterile\" from the manufacturers. Proper aseptic techniques were practiced and used accordingly. HEp-2 cells HEp-2 cells, a transformed human cell line isolated from epidermoid carcinoma of the larynx, were purchased from American Type Culture Collection (ATCC) (Manassas, V A ) . HEp-2 cells were seeded in T75 flasks (Corning; Corning, N Y ) containing minimum essential medium (MEM) (Life Technologies, Burlington, O N , Canada) supplemented with 10% heat-inactivated fetal bovine serum (FBS; Life Technologies) and 50 ug/mL gentamycin (Life Technologies) and incubated in an incubator (Fisher Scientific, Nepean, ON) at 37\u00C2\u00B0C in an atmosphere of humidified air and 5% CO2. The cell line was maintained by sub-culturing the confluent monolayers. When cells were confluent the culture medium was removed and the monolayer was rinsed with sterile phosphate buffered saline (PBS, pH 7.4; Sigma). A 1 mL solution of trypsin (0.25%)-EDTA (0.02%), prewarmed at 37\u00C2\u00B0C, was added to the flask. The flask was rocked gently to ensure even distribution of the trypsin solution and incubated at 37\u00C2\u00B0C for 5 minutes. The cells were detached by slapping the flask in a vertical position. The enzymatic action of trypsin was inhibited by the addition of fresh culture medium containing FBS. The sub-cultivation ratio as recommended by the supplier for HEp-2 cells is between 1:4 and 1:10. Freezer stocks of HEp-2 cells were prepared by the above trypsinization procedure followed by a 30 5-minute centrifugation in PBS at 1000 x g. The cells were re-suspended in a 3 mL solution of culture medium: DMSO (9:1) and stored as 1 mL aliquots in cryovials. Cells were frozen in a \"slow-freeze\" procedure by standing on ice for a half-hour, kept in the -20\u00C2\u00B0C freezer for 1 hour, then kept in the -70\u00C2\u00B0C freezer overnight and transferred to liquid nitrogen tanks for long term storage the following day. L929 Fibroblasts TNFcc-sensitive L929 fibroblasts were a gift from Dr. R.R. Schellenberg, McDonald Research Laboratories, St. Paul's Hospital, Vancouver, BC. L929 cells were cultured in T75 flasks of RPMI medium (Life Technologies) containing 10% FBS, 10 m M Glutamine, 100 units/mL penicillin, 100 |ig/mL streptomycin and 250 ng/mL amphotericin B. The sub-cultivation ratio as recommended by the original supplier (ATCC) is 1:6. The procedure to maintain the cell line and to store freezer stocks is as described for HEp-2 cells. VIRUS CULTURE R S V Propagation The Long strain, type A of human RSV was purchased from A T C C . RSV stocks were propagated on monolayers of HEp-2 cells in M E M supplemented with 2% FBS and 50 jig/mL gentamycin. RSV infection of HEp-2 cells induces the cytopathic effect of multiple cells fusing to form a mega-cell or syncytium. RSV was harvested when syncytial formation peaked in the monolayers of HEp-2 cells. A crude RSV stock was prepared by scraping infected cell monolayers with a sterile cell scraper and the cells and supernatant were transferred to 50 mL tubes for short term storage in the -70\u00C2\u00B0C freezer. This crude preparation of R S V was used for regeneration of viral stocks. 31 Harvest and Concentration of RSV Concentrated stocks of RSV were prepared by mechanical disruption of the syncytia-filled monolayers of HEp-2 cells using sterile 3 mm glass beads (Fisher Scientific) over a vortex for 30 seconds. The cell suspension was transferred to 50 mL tubes and underwent centrifugation at 1500 x g, 4\u00C2\u00B0C for 15 minutes to sediment the cellular debris. The clear supernatant (containing free virus and soluble macromolecules such as inflammatory mediators) was applied on to Centriplus\u00E2\u0084\u00A2 concentrators (Amicon, Beverly, M A ) with molecular cut-off at 100,000 Daltons to concentrate the virus as well as to remove soluble macromolecules synthesized during viral infection of HEp-2 cells. The clear supernatant was spun in the concentrator unit at 3000 x g, 25\u00C2\u00B0C for 75 minutes and the virus-enriched retentate was stored in 1 mL aliquots in the -70\u00C2\u00B0C freezer. RSV titer determination HEp-2 cells were seeded in 6-well plates (Corning; Corning, N Y ) at a density of 5 x 105 cells/well. Overnight culture at this density usually results in cells reaching approximately 80-90% confluency. Ten-fold serial dilutions (10\"1 to 10\"7) of RSV stocks were prepared using M E M . HEp-2 cells were exposed to 1 mL of the serially diluted RSV and viral adsorption was allowed to take place over 90 minutes at 37\u00C2\u00B0C in a 5% CO2 incubator (Fisher Scientific). The plates were gently agitated intermittently at 15-minute intervals to ensure even distribution of virus over the cells. The wells were washed in PBS and 1 mL of medium-agarose mixture (1:1, 2X M E M with 2% FBS:1% agarose) was applied. The agarose mixture was allowed to solidify at room temperature before returning to the CO2 incubator. The cells were cultured until syncytia were visible ( 7 - 1 0 days) and were fixed in 4% paraformaldehyde (Ted Pella Inc, Redding, CA) at room temperature for 30 minutes. The agarose layers were removed and the wells rinsed with dH20. The adherent monolayer was fixed in methanol (BDH Chemicals, 32 Toronto, ON, Canada) and air-dried. To enhance visualization of the syncytia formation, the cells were stained with 0.1% neutral red (Life Technologies) for 1 minute and then washed with running tap water. The syncytia were counted under an inverted light microscope. The number of plaque forming units per milliliter (pfu/mL) (i.e., the amount of replicating virus per milliliter of stock) was determined by the number of syncytia counted divided by the corresponding dilution factor. ANIMALS Juvenile Cam Hartley guinea pigs of both sexes, 22 to 29 days old (250 g to 300 g body weight) and female retired breeders (>1000 g) were purchased from Charles River Laboratories (Montreal, QC, Canada). The animals were housed in metal cages with corncob bedding and access to guinea pig chow (Purina, Ralston Purina Company, St. Louis, MO), alfalfa cubes and water. The animals were allowed to acclimatize for at least 5 days before being used for experiments. The guinea pigs were maintained in accordance with standards of the Canadian Council on Animal Care (197). A L V E O L A R MACROPHAGE HARVEST Bronchoalveolar Lavage The guinea pigs were killed by intraperitoneal administration of pentobarbital (Euthanyl; M T C Pharmaceuticals; Cambridge, O N , Canada) at a dose of 40mg/kg body mass. When the animal was fully euthanized (no response to forceps pinch test or spontaneous respiration), bronchoalveolar lavage was performed in situ. The animal was positioned ventral side up and a longitudinal cut was made through the skin from below the jaw line to the bottom of the sternum. The trachea was exposed and a small horizontal cut was made between two cartilage rings. A 21-gauge cannula was inserted approximately 1 cm into the trachea and secured with a piece of 33 suture thread. The lungs were lavaged by intratracheal instillation of 5 mL aliquots of sterile non-pyrogenic normal saline solution (Baxter, Toronto, ON, Canada), pre-warmed at 37\u00C2\u00B0C, totaling 100 mL. The instilled fractions were gently aspirated and pooled together in 2 x 50 mL centrifuge tubes and kept on ice until processed. This procedure of bronchoalveolar lavage performed on guinea pigs results in a typical yield of 10-15 million A M per animal. Cell Processing The B A L fluids were centrifuged at 500 x g for 10 minutes at 4\u00C2\u00B0C. The resultant cell pellets from the same animal were pooled together and re-suspended in 5 mL Hank's balanced salt solution (HBSS, Life Technologies) and counted in a hematocytometer. Isolation of Macrophage Subpopulations Columns of discontinuous gradients were prepared fresh in 15 mL transparent polypropylene centrifugation tubes (Corning) immediately before use. This method was adapted from Dakhama et al. (83) using 3 concentrations of metrizamide (Nycomed Pharma AS; Oslo, Norway), 18%, 20% and 22% prepared in H-saline2. Cell pellets from the same animal were pooled together and counted; the cell suspensions were spun down again at 500 x g for 10 minutes at 4\u00C2\u00B0C and the pellet re-suspended in 3 mL of 18% metrizamide solution. This cell preparation was gently overlaid a 15-mL polypropylene centrifugation tube containing 3 mL of 20% metrizamide overlaid to another 3 mL of 22% metrizamide solution. A 4 mL overlay of H -saline was applied over the cell suspension to prevent dehydration during centrifugation. After centrifugation at 1,200 x g for 45 minutes at 18\u00C2\u00B0C, the cells were collected from each interface and washed in 5 mL HBSS. 2 H-saline is a HEPES-buffered saline solution containing lOraM HEPES (pH 7.4, Sigma-Aldrich, Toronto, ON, Canada), 150mM NaCl, 0.1% glucose and 0.1% gelatin. 34 In vitro RSV exposure After washing, viability of the A M was determined by trypan blue dye exclusion test and the cell concentration was adjusted to 1 x 106/mL. The A M were seeded in 6-well plates at a density of 2 x 106 cells/well and allowed to adhere for 30-60 minutes. For an infection with a multiplicity of infection (m.o.i.) of 3, the viral stock concentration was adjusted to 6 x 106 pfu/mL of which 1 mL was applied to each well. Viral adsorption with intermittent agitation of 6-well plates was carried out over 90 minutes. The wells were washed with PBS and fresh M E M containing 5% FBS were added (2 mL/well) followed by overnight incubation at 37\u00C2\u00B0C in a 5% CO2 incubator. 35 C H A P T E R 5 E F F E C T O F C E L L M A T U R A T I O N - A N E M S T U D Y INTRODUCTION A M exist as an heterogeneous population in the lung in vivo (143). Using density gradient centrifugation, subpopulations of A M can be separated based on the inverse relationship of cellular maturation and buoyant density (141, 143). By examining subpopulations of A M based on their stage of maturation, our laboratory (83) has demonstrated that cell maturation may be an important factor in the susceptibility of these cells to acute RSV infection in vitro. Using immunocytochemistry and viral plaque assay, to determine the proportion of RSV-positive cells and the amount of intracellular replicating virus respectively, Dakhama et al. (83) showed that immature A M were significantly more susceptible to RSV infection and supported RSV replication more efficiently compared to their matured cellular counterparts. While it is known that A M of different maturation stages are morphologically and functionally different (141,143), it is not understood if the different subpopulations respond differently to RSV infection. Dakhama et al. (83) have shown that immunostaining of RSV antigens manifest a distinct granular pattern in the cytoplasm of mature A M while a less granular and more diffuse cytoplasmic staining is found in intermediate and immature A M . These observations suggest that mature A M may be structurally better equipped to deal with RSV infection when compared to the relatively immature A M . Ultrastructural studies of RSV infection in HEp-2 cells (18, 22) and eosinophils (109) have been documented; however, prior to the publication of this work (83), there have been no ultrastructural studies of RSV in A M . Arslanagic et al. (18) showed two mechanisms by which RSV mature in HEp-2 cells. Maturation of RSV particles occurred on the internal vesicle membrane within the cytoplasm before being delivered to the plasma membrane by transport vesicle and were subsequently released into the extracellular space by exocytosis. 36 The other pathway for RSV maturation, a widely accepted maturation process for most paramyxoviruses, newly synthesized viral proteins are assembled and packaged on the plasma membrane of infected cells and mature by budding through the plasma membrane (13). In the work by Garcia et al. (22), RSV antigens, as recognized by RSV-specific antibodies conjugated to colloidal gold particles, were detected in electron-dense inclusion bodies found in the cytoplasm in close proximity to cell nuclei of RSV-infected HEp-2 cells. Kimpen et al. (109) demonstrated that RSV particles could be taken up by phagocytic vacuoles in eosinophils within 2 hours following exposure of the cells to virus. HYPOTHESIS AND SPECIFIC AIMS On the basis of the above observations, we speculate that the maturation stage of A M may be a determining factor in the outcome of RSV infection and the Working Hypothesis is: Differences in RSV replicative outcome in subpopulations of guinea pig A M may be due to differences in the intracellular compartmentalization of RSV during its life cycle in these cell subpopulations. The Specific Aims of this study are: 1. To isolate heterogeneous guinea pig A M into 3 subpopulations by density gradient centrifugation. 2. To determine intracellular localization of RSV antigens in these A M subpopulations following in vitro infection using immunogold labeling and transmission electron microscopy. 37 EXPERIMENTAL PROTOCOLS Study Design The design of this study is shown in Figure 5.1. Briefly, juvenile female guinea pigs (n=3) were killed and heterogeneous A M were obtained by in situ bronchoalveolar lavage. The A M were isolated into three subpopulations, \"immature\", \"intermediate\" and \"mature\", by density gradient centrifugation. The cells were transferred to 6-well plates (2x106 cells per well) after washing in cold RPMI medium and left in the incubator (37\u00C2\u00B0C, 5% CO2) for 30 minutes to allow viable A M to adhere to the plates. Non-adherent cells were washed off and the 3 subpopulations of A M were exposed to RSV at an m.o.i. of 3. At 24 hours post-infection, the cells were collected into pellets and processed by standard E M procedures before being embedded and cured in plastic blocks. Ultra-thin sections of cells were prepared and immunogold labeling was performed. Examination of these cells was carried out by transmission electron microscopy. Positive control HEp-2 cells were used in parallel with the guinea pig A M through the experimental procedures. 38 Broncho-Alveolar Lavage Density Gradient Centrifugation ^immature^ intermediate^ AM /^Iviature^ RSV infection (moi =3) < Immunogold Labeling & TEM Figure 5.1 Experimental design of E M study-infection in guinea pig A M . a cell maturation effect on in vitro RSV 39 Immunogold Labeling & Transmission electron microscopy Cell Processing and Fixation At 24 hours post-infection, RSV-exposed A M were scraped off 6-well plates with a sterile cell scraper and washed in PBS. The A M were collected into a pellet by centrifugation at 4\u00C2\u00B0C, 400 x g for 10 minutes. After aspirating off the PBS, the cell pellets were gently dislodged and fixed in 0.1 M sodium cacodylate buffer (pH7.4, Marivac Ltd., Halifax, NS, Canada) containing 2.5% glutaraldehyde (BDH Chemicals) for 1 hour at 4\u00C2\u00B0C. The cells were spun down and resuspended in a minimum amount of cacodylate buffer. This cell suspension was transferred to rubber mould. A n equal volume of tepid 2% (w/v) low-melting point agarose (Sigma) was added to the cell suspension, mixed thoroughly and allowed to set in the refrigerator. The agar block was cut into 2 mm 3 cubes, transferred to glass vials, washed thrice in cacodylate buffer and then fixed in 1 % osmium tetroxide for 1 hour in the dark. The cubes were washed in 3 changes of dH20 followed by an hour fixation in saturated uranyl acetate. Excess uranyl acetate was washed off in 3 changes of dHzO. Unless otherwise mentioned, all electron microscopy products/reagents were purchased from Electron Microscopy Services, Fort Washington, PA. Infiltration and Embedding Prior to infiltration, the cubes containing A M were passed through a series of increasing concentrations of alcohol, (30%, 50%, 70%, 90% and 3X100%) at 10-minute intervals. The cubes were washed thrice in 100% propylene oxide for 10 minutes each. Immediately before use, equal volumes of Epon Mixture I was mixed with propylene oxide and used as infiltration fluid. After aspirating off the propylene, the freshly prepared infiltrating fluid was added to the cubes. The vials of cell cubes were left to infiltrate overnight on a rotator with their lids off. The 3 Recipe for Epon Mixture I: Prepare Stock using 62 mL of Epon 812, 81 mL of Araldite 502 and 5 mL of DBP (Dibutyl Phthalate). Mix 1 part Stock to 1 part DDSA (Dodecenyl Succinic Anhydride) and 1.5% (v/v) DMP30 (2,4,6-Tri (dimethylaminomethyl) phenol). 40 cubes were then transferred the following day to rubber moulds, embedded in fresh 100% Epon Mixture I and left to cure overnight in a 60\u00C2\u00B0C oven in a fume hood. Grid Preparation and Cutting Specimen grids made of nickel with a diameter of 3.05 mm and standard hexagonal mesh were used for supporting ultra-thin sections of cells. The grids were coated with 0.5% formvar in ethylene dichloride. Semi-thin sections (0.5 |j.m thickness) were cut using a glass knife fitted with a trough of dF^O. These sections were stained with Toluidine Blue O and examined under light microscope to locate areas of interest. Ultra-thin sections, between 60 nm (gray color) and 90 nm (silver color) in thickness, were sectioned using a fresh glass knife. The sections floating in the trough of dthO were transferred onto the dull side of formvar-coated grids with the aid of a wire loop. Immunogold Labeling and Visualization Ultra-thin sections of specimens that were mounted on nickel grids were pretreated with 2% sodium metaperiodate to unmask antigenic sites (198). The sections were first incubated with 5% normal goat serum to minimize background staining and then incubated with N C L -RSV2 (Novocastra Laboratories; Newcastle-Upon-Tyne, UK) , a pool of monoclonal antibodies with specificities for RSV fusion protein (F), phosphoprotein (P) and the 22K membrane-associated protein (M2). The feasibility of the antibodies directed against the P and M2 proteins have been demonstrated by Garcia et al. (22) for intracellular localization. The primary antibody was used at a empirically determined working dilution of 1:50 in PBS for 1 hour at 37\u00C2\u00B0C. Consecutive sections incubated with non-immune mouse IgG] served as controls for immunostaining. Excess primary antibodies were washed off using PBS. Specific antibody-antigen reaction sites were labeled by incubation of goat anti-mouse immunoglobulins conjugated to 10 nm diameter gold particles (1:50 dilution, Dakopatts) for 30 minutes at 25\u00C2\u00B0C. Three changes of PBS plus a final wash in dF^O was performed. The sections were 41 counterstained with uranyl acetate and lead citrate prior to viewing under a Philips 400 electron microscope (Philips 400; N . V . Philips' Gloeilampen-fabrieken; Eindhoven, The Netherlands). RESULTS Isolation of A M subpopulations Figure 5.2 shows heterogeneous A M harvested by bronchoalveolar lavage isolated into 3 distinct fractions on a metrizamide gradient following density gradient centrifugation. The subpopulation of A M recovered at the 20-22% metrizamide interface was designated as \"immature\", those recovered at the 18-20% metrizamide interface were designated as \"intermediate\", and those residing in the 18% metrizamide-H-saline interface were \"mature\" A M (83). Localization of R S V particles A l l three subpopulations of A M were exposed to RSV (m.o.i =3) and immunolabeled with a pool of anti-RSV monoclonal antibodies. Examination by transmission electron microscopy revealed aggregates of colloidal gold particles in the lysosomal compartment of mature A M (Figure 5.3A). In contrast, clusters of gold particles were seen in perinuclear distribution in the cytoplasm of intermediate A M (Figure 5.3B) and in immature A M (Figure 5.3C), the gold particles associated with electron-dense material, that had the ultrastructural appearance of free RSV nucleocapsid. Positive control RSV-infected HEp-2 cells (Figure 5.4), which showed gold particles within cytoplasmic regions and budding of virus progeny around the cell membrane, confirmed the specificity of immunogold labeling by our T E M technique. 42 Figure 5.2 Isolation of A M subpopulations by density gradient centrifugation. Heterogeneous A M obtained from bronchoalveolar lavage were separated into hypodense mature (A), intermediate (B) and high density immature (C) subpopulations on a discontinuous metrizamide gradient. 43 r~n Figure 5.3 Electron micrograph of immunogold labeling of RSV particles using anti-RSV antibodies in A M subpopulations. Scale bar represents 50 nm. (A) Mature A M : colloidal gold labeled RSV particles (arrow) were observed in phagolysosome (pl). (B) Intermediate A M : colloidal gold particles (arrow) found in cytoplasm near the nucleus (N). (C) Immature A M : colloidal gold particles localized to viral nucleocapsid (asterisk) in cytoplasm. (D) Immature A M : negative control. Figure 5.4 Electron micrograph of RSV-infected HEp-2 cell. Arrow indicates colloidal gold localized to RSV progeny budding from cell membrane. Scale bar represents 100 nm. 45 DISCUSSION The objectives of this study were to localize RSV proteins within infected A M by electron microscopy and compare intracellular compartmentalization of RSV in the three subpopulations of guinea pig A M . The results showed that immunogold labeled-RSV particles were found in different intracellular compartments of A M according to their degree of maturation. More specifically, colloidal gold particles were observed in the lysosomal compartment of mature A M , whereas gold particles were found in the cytoplasm of intermediate and immature A M . These observations are consistent with and extend the light microscopic observations made by Dakhama et al. (83) in that RSV infection of mature A M yielded a granular staining pattern (which we interpreted as corresponding to lysosomes) while that of intermediate and immature A M produced a diffuse cytoplasmic staining pattern (which we interpreted as corresponding to free cytosolic virus). Ultrastructural studies on functional studies of subpopulations of human A M have been described (199). Transmission electron microscopic analysis by Nakstad et al. (199) revealed mostly primary lysosomes and very few secondary lysosomes in high density immature A M in contrast to large mature A M with numerous secondary lysosomes. Primary lysosomes are membrane-bound organelles which contain lytic enzymes for intracellular digestion but have not yet become engaged in enzymatic digestive activities. When a phagosome (with its phagocytosed material) fuses with a primary lysosome, the resulting merged organelle is referred to as a secondary lysosome or phagolysosome (156). Thus localization of colloidal gold particles in the phagolysosomes of mature A M , in contrast to the free cytoplasmic colloidal gold particles found in intermediate and immature A M , suggests that mature A M were more able to contain the viral particles within the lysosomal compartment than less mature A M subpopulations. Since virus-associated colloidal gold particles were found in the phagolysosomes of mature A M , this localization may be pertinent to the mechanisms by which 46 mature A M inactivate RSV and restrict viral replication. On the other hand, the observation of virus-associated colloidal gold particles free in the cytoplasm of intermediate and immature A M is similar to observations made in RSV-infected epithelial cells (22). Furthermore, in immature A M , the colloidal gold particles were associated with electron dense material that showed morphologic features typical of paramyxovirus nucleocapsids (22, 200). The observations made in intermediate and immature A M suggest that these cells were less able to inactivate RSV and thereby have a tendency to support viral replication. Qualitative observations, such as those described here, represent a limitation of the current study but are nonetheless valuable because the differences in R S V - A M interactions show distinct patterns of ultrastructural localization that extend light microscopic observations. Although the proportions of A M subpopulations in children have not been determined (due to practical and ethical reasons), most of the A M in the lungs of normal adults are of the mature subpopulation which are relatively more capable of conferring protection (201); this may be an explanation for immunocompetent adults to rarely develop serious RSV-bronchiolitis. Investigations to examine the relative proportions of A M subpopulations between the ages in animal studies (see Chapter 6 - Effect of age of host animal) would provide new information to improve our understanding of the interactions between RSV and A M . In summary, the ultrastructural observations of immunogold labeled-RSV particles in different intracellular compartments of A M according to their degree of maturation supports the hypothesis that differences in RSV outcome in subpopulations of A M may be due to differences in intracellular compartmentalization of RSV, with the mature subpopulation of A M containing viral antigens within phagolysosomes. 47 C H A P T E R 6 E F F E C T O F A G E O F H O S T A N I M A L O N R S V I N F E C T I O N INTRODUCTION The age of the host animal is a major determinant in the ability of the cell to contain viral infections. Increased resistance to disease in the course of growth, development and maturation of the host appears to be a general characteristic in humans and other species. Age-dependent resistance of macrophages to viral infections has been demonstrated in animal models. Some examples, included but not limited to, are herpes simplex virus (HSV) (202, 203) and rabies virus (204). Several theories to explain the age-related resistance against viral infections have been postulated; included are maturation of immunological reactivity, augmented IFN production and reduction in number of cell surface viral receptors (205). As mentioned in the overview, the age of the host is one of the determinants affecting RSV infection. In children, the severity of primary pulmonary disease due to RSV infection is inversely proportional to age (1). The nature of the maturation process that renders in most adults the ability to confine RSV infection in the upper respiratory tract is still not well understood. Although numerous animal models of RSV infection have been developed, only a limited number of these models have included the age dependency phenomenon (73). Using the ferret model, Prince and Porter (206) showed that RSV could replicate in the lungs of infant but not adult animals. Similarly, data from our laboratory has demonstrated that juvenile guinea pigs are more susceptible to acute bronchiolitis than adult guinea pigs when inoculated with human RSV (207). While the effect of age on in vitro RSV infection in subpopulations of A M has not been examined, it has been studied in different virus-host systems. Hirsch et al. (202) showed that HSV-infected macrophages from suckling mice released more progeny virus than adult mouse 48 macrophages. In vitro studies by Morgensen (203) showed that HSV-2 infection of peritoneal macrophages from 3-week old and 8-week old mice correlated age-related resistance with an increased restriction of viral replication, as determined by plaque assay. Recent studies of in vitro RSV infection in A M have demonstrated that viral infection is restricted to a subset or subpopulation of cells (83, 108, 113). Midulla et al. (108) studied the \"permissiveness\" of various cell types to RSV infection by examining the proportion of RSV-positive cells following in vitro exposure to RSV. Cirino et al. (113) showed that only approximately one-third of A M exposed to RSV were capable of replicating RSV and restriction of viral replication may be related to cellular differentiation. Dakhama et al. (83) extended this observation: in addition to examining the proportion of RSV-positive cells following in vitro RSV infection, they included in their investigation the amounts of replicating virus in RSV-exposed A M . However, none of these studies examined the effect of age of the host animal on RSV infection in conjunction with subpopulations of A M . Various investigators (83, 108, 113) have used the term \"permissiveness\" in their reports in a rather vague manner and without precise definition. To avoid confusion, this term will not be used in this thesis. A more complete understanding of the interaction between RSV and A M could be achieved by considering the ability of RSV to enter cells as well as the ability of cells to support viral replication as one entity. Therefore, in this thesis, a new term, RSV Yield, will be used. The RSV Yield of a cell is defined as the amount of viral replication (as determined by plaque assay) per RSV-immunopositive cell (as determined by RSV immunostaining). Mathematically, R S V Yield = (pfu per million A M ) + (% RSV-immunopositive A M ) The investigations in this study examined the effect of age of guinea pigs on in vitro RSV infection by analyzing the distribution of the A M subpopulations and the RSV Yield of A M subpopulations between juvenile and adult animals. 49 HYPOTHESIS AND SPECIFIC AIMS The Working Hypothesis of this study is: The susceptibility of guinea pig A M to in vitro RSV infection is dependent on the age of the host animal and the state of cell maturation. The Specific Aims of this study are: 1. To examine the distribution of AM subpopulations between juvenile and adult guinea pigs. 2. To determine and compare the uptake of RSV in A M subpopulations isolated from juvenile and adult guinea pigs. 3. To determine and compare RSV replication of RSV in AM subpopulations isolated from juvenile and adult guinea pigs. 4. To determine and compare the RSV Yield in AM subpopulations isolated from juvenile and adult guinea pigs. EXPERIMENTAL PROTOCOLS Study Design The design of this study is shown in Figure 6.1. Six adult female guinea pigs (retired breeders, >1000 g) and five one month-old, juvenile female guinea pigs (250 - 300 g) were used in this study. Since puberty in the female guinea pigs occur around 10 weeks of age, young sows at one month of age are considered to be sexually immature (208). Heterogeneous AM were obtained by bronchoalveolar lavage immediately following killing. The cells were isolated into 3 subpopulations (immature, intermediate and mature) by means of density gradient centrifugation. These cells were exposed to RSV at a m.o.i.=3 for 90 minutes with intermittent agitation, and harvested at 24 hours post-exposure. The susceptibility of A M subpopulations to 50 RSV infection and replication were determined by a combination of immunostaining of cytospin preparations and viral plaque assays, and RSV Yield was calculated. RSV infection (mo.i=3) I 1. Immunocytochemistry 2. Plaque Assay Figure 6.1 Experimental design for study of effect of age of host animal on in vitro RSV infection in guinea pig A M . 51 Cytospin preparation RSV-exposed A M and control cells (HEp-2 cells) were collected from 6-well plates by scraping off the monolayers using a sterile cell scraper and transferred to 15 mL polypropylene tubes. The cells were collected by centrifugation at 200 x g for 10 minutes in a bench top centrifuge. The resultant cell pellets were re-suspended in M E M supplemented media plus 1% FBS. The cell density was adjusted to 0.5 x 106 cells/mL. 100 \iL of each sample was applied to the cytospin chamber that had been assembled with a blotter (Shandon Inc, Pittsburgh, PA) and a labeled adhesion microscope slide (Histobond\u00C2\u00AE, Marienfeld, Germany). At this cell density, a target of 50,000 to 100,000 cells per cytospin spot was obtained. The assembled cytospin chambers were spun at 200 x g for 5 minutes at room temperature. The prepared slides were air-dried and fixed in methanol (BDH Chemicals) for 5 minutes. I mmunocy tochemistry RSV Immunostaining RSV antigens in cytospin preparations were detected by immunostaining using RSV mouse monoclonal antibody (NCL-RSV2, Novocastra Laboratories, New Castle-upon-Tyne, UK) in conjunction with the Vectastain\u00C2\u00AE Avidin-Biotin Complex-alkaline phosphatase (ABC-AP) kit (Vector Laboratories, Burlingame, CA). NCL-RSV2 is made up of a pool of monoclonal antibodies with specificities for RSV fusion protein, phosphoprotein and M2 protein. The staining protocol as recommended by the manufacturers has been modified for cytospin preparations as follows. Slides were boiled in 0.01 M sodium citrate buffer (pH 6.0, B D H Chemicals) for 10 minutes, cooled and rinsed in dH20. The slides were incubated for 20 minutes with diluted normal blocking serum4. Excess serum was blotted off, and the NCL-RSV2 antibody solution at 10 Hg/mL was applied to each section and incubated for one hour. The 4 Normal blocking serum: 3 drops of stock to 10 mL TBS (10 mM sodium phosphate, pH 7.5, 0.9% saline). 52 slides were washed thrice in TBS and incubated with the biotinylated secondary antibody5 for 30 minutes. After another 3 x 5 minute washes, the slides were incubated with VECTASTAIN\u00C2\u00AE A B C - A P Reagent6 for 30 minutes. Following that, the slides were washed again and incubated with Vector\u00C2\u00AE Red Alkaline Phosphatase Substrate7 (Vector Laboratories) for 30 minutes. A final wash in tap water was carried out before mounting in VECTASHIELD\u00C2\u00AE mounting medium and viewed under light microscope. For every cytospin sample tested, control slides using IgG] of equivalent protein concentration to substitute the primary antibody were carried out in parallel. A l l incubations were performed at room temperature. Scoring and Statistical Analysis A total of 300 A M were scored and the proportion of RSV-positive cells were reported as a percentage of cells counted. The data were reported as the means of each group \u00C2\u00B1 SEM, analyzed using A N O V A and the Bonferroni procedure was used to correct for multiple comparisons. A two-tailed p value of < 0.05 was considered as statistically significant. Viral Plaque Assay Quantification of RSV progeny titer After the A M had been exposed to RSV for 90 minutes at an m.o.i. of 3 and allowed to incubate overnight, the cells were gently scraped with sterile cell scraper (Fisher Scientific) and the cell suspensions collected into 2 mL sterile eppendorf tubes. Sterile glass beads (Fisher Scientific) were added and the tubes were vortexed for 5 minutes to release cell-associated virus. Ten-fold serial dilutions (10\" to 10\") of each sample were prepared using culture medium. 1 mL of each dilution was applied onto a monolayer of HEp-2 cells in 6-well plates. A l l samples 5 Biotinylated secondary antibody: working concentration is 1 drop of stock in 10 mL TBS. 6 ABC-AP Reagent: 2 drops of Reagent A + 2 drops of Reagent B in 10 mL TBS. Mix and stand for 30 minutes before use. 7 Vector Red Substrate: Working solution is prepared immediately before use - 2 drops of Reagent 1 + 2 drops of Reagent 2 + 2 drops of Reagent 3 in 5 mL of 100 mM Tris-HCl buffer, pH 8.2-8.5 53 were plated in duplicate. Subsequent steps for this assay involved washing of virus-exposed cells with PBS, agarose overlay followed by incubation over a period of 7 to 10 days of incubation and syncytia scoring. Full details of these methods are described in the section for RSV titer determination (Chapter 4, General Materials & Methods). Scoring and Statistical Analysis The number of syncytia formed in the different serial dilutions of sample applied was scored. The dilution, which yields 25-75 syncytia, was used to determine the number of plaque forming units (pfu) per 106 cells. The data, reported as the means \u00C2\u00B1 S E M , were analyzed using A N O V A and the Bonferroni method was used to correct for multiple comparisons. A value of p<0.05 (two-tailed) was considered as statistically significant. 54 RESULTS Distribution of A M subpopulations The distribution of A M subpopulations from juvenile and adult guinea pigs was examined and is presented in Table 6.1. There were no significant differences in the distribution of A M subpopulations between adult and juvenile guinea pigs. Table 6.1 Distribution of A M subpopulations in adult and juvenile guinea pigs. Adult .1 uvenile Mature A M 38.8 \u00C2\u00B15 .9 46.6 \u00C2\u00B1 1.9 Intermediate A M 31.5 \u00C2\u00B11.4 29.8 \u00C2\u00B14 .4 Immature A M 29.7 \u00C2\u00B1 5.0 | 23.6 \u00C2\u00B14.8 Data represent the proportion (mean \u00C2\u00B1 SEM) of mature, intermediate and immature A M . 55 Effect of Age on RSV Immunopositivity Figure 6.2 shows the staining pattern of RSV-infected subpopulations of A M in comparison to the IgGi non-specific staining control. RSV antigens were localized in the cytoplasm of RSV-infected A M . Control slides incubated with immunoglobulin (IgGi) did not exhibit any immunostaining. Figure 6.3 shows results of the percentage of RSV-immunopositive cells, determined by scoring at least 300 cells per specimen under the light microscope. The proportion of RSV-immunopositive cells in juvenile guinea pigs are significantly higher compared to their respective maturation stages in adult guinea pigs (p<0.0001). In juvenile animals, the percentage RSV-immunopositive A M from the mature subpopulation was significantly less than those in the intermediate (mature vs. intermediate (mean \u00C2\u00B1 SEM): 10.6 \u00C2\u00B1 0.7% vs. 18.6 \u00C2\u00B1 0.9%, p<0.0002) as well as immature (mature vs. immature: 10.6 \u00C2\u00B1 0.7% vs. 20.6 \u00C2\u00B1 1.3%, p<0.002) subpopulations. There were no significant differences in the proportion of RSV-immunopositive cells between intermediate and immature A M subpopulations from juvenile animals. In contrast, there were no significant differences among the three subpopulations of A M from adult guinea pigs. 56 up?'. \u00E2\u0080\u00A2 \u00E2\u0080\u00A2* \u00E2\u0080\u00A2 * c D \u00E2\u0080\u00A2 # Figure 6.2 Immunostaining of mature (A), intermediate (B), and immature (C) subpopulations of guinea pig A M with anti-RSV antibody. The negative control (D) was an irrelevant isotype-matched IgGi. Hematoxylin counterstained. Scale bar represents 50 um. Figure 6.3 Effect of age of host animal on RSV immunopositivity. Data represent percentage of RSV immunopositive A M (mean \u00C2\u00B1 SEM) in mature (white bars), intermediate (gray bars) and immature (black bars) subpopulations from adult and juvenile guinea pigs. 'a' pO.OOOl adult vs. juvenile guinea pigs in respective A M subpopulations. 'b' p<0.0002 mature A M vs. intermediate A M in juvenile guinea pigs, 'c ' p<0.002 mature A M vs. immature A M in juvenile guinea pigs. 57 Effect of Age on RSV Replication Infectious, replicating R S V within A M was released by mechanical disruption with sterile glass beads. Ten-fold serial dilutions of the virus were used to infect HEp-2 cells which form large syncytia when exposed to RSV (Figure 6.4). The number of R S V progeny from A M was determined as the product of the number of syncytia observed and the reciprocal of the dilution factor in which the syncytia were scored. Figure 6.5 shows the RSV progeny in different A M subpopulations from adult and juvenile guinea pigs. The RSV progeny in all three subpopulations of A M from juvenile guinea pigs was significantly greater than their respective counterparts from adult animals (p<0.002). As shown in Figure 6.5, the immature subpopulation of A M from juvenile animals had significantly more R S V progeny compared to the intermediate ((mean \u00C2\u00B1 SEM) 4554 \u00C2\u00B1 7 5 1 vs. 975 + 257, p<0.006) and mature (4554 \u00C2\u00B1 751 vs. 252 \u00C2\u00B1 27, p<0.003) subpopulations. In addition, intermediate A M contained more RSV progeny than the mature A M in juvenile animals (975 \u00C2\u00B1 257 vs. 252 \u00C2\u00B1 27, p<0.03). In contrast, there were no significant differences in the RSV progeny from the three A M subpopulations isolated from adult animals. 58 I Figure 6.4 Formation of syncytium as a result of infection and fusion with neighboring cells on monolayer of HEp-2 cells stained with neutral red (0.1 %). Scale bar represents 50 um. Figure 6.5 Effect of age of host animal on RSV replication. Data represent RSV progeny per million A M exposed to RSV in mature (white bars), intermediate (grey bars) and immature (black bars) subpopulations from adult and juvenile guinea pigs. 'a' p<0.002 adult vs. juvenile guinea pigs in respective subpopulations. 'b' p<0.006 immature A M vs. intermediate A M from juvenile guinea pigs, 'c ' p<0.003 immature A M vs. mature A M from juvenile guinea pigs, 'd ' p<0.03 intermediate A M vs. mature A M from juvenile guinea pigs. 59 Effect of Age on RSV Yield The RSV Yield in A M from adult and juvenile guinea pigs is presented in Table 6.2. The data show that RSV Yield of immature A M from juvenile guinea pigs was 5 times greater than that from adult guinea pigs (Immature A M , juvenile vs. adult: 228 \u00C2\u00B1 45 vs. 45 \u00C2\u00B1 13, p<0.002). Comparison of RSV Yield on intermediate A M between adult and juvenile did not reveal significant differences and there was a trend (p=0.08) for higher RSV Yield in mature A M from adult versus juvenile animals. In juvenile guinea pigs, the data indicate that RSV Yield of immature A M was significantly higher than the intermediate (228 \u00C2\u00B1 45 vs. 52 \u00C2\u00B1 13, p<0.008) or mature (228 \u00C2\u00B1 45 vs. 24 + 3, p<0.008) A M subpopulations. In contrast, there were no significant differences in RSV Yield among the A M subpopulations from adult guinea pigs. Table 6.2 RSV Yield in subpopulations of A M from adult and juvenile female guinea pigs. Adult .1 uvcnilc Mature A M 40 \u00C2\u00B1 9 24 \u00C2\u00B1 3C Intermediate A M 86 \u00C2\u00B1 5 8 5 2 \u00C2\u00B1 1 3 b Immature A M 45 \u00C2\u00B1 1 3 2 2 8 \u00C2\u00B1 4 5 a Data represent the RSV Yield (mean \u00C2\u00B1 SEM) of subpopulations of guinea pig A M . 'a ' p<0.002, Immature A M : adult vs. juvenile animals. 'b' p<0.01, Juvenile animals: immature A M vs. intermediate A M . 'c ' p O . O l , Juvenile animals: immature A M vs. mature A M . 60 Summary of Results In summary, the data in this study showed that there were no differences in the distribution of A M subpopulations between adult and juvenile guinea pigs. However, in juvenile animals, the uptake of RSV by immature A M and the corresponding R S V progeny titer was significantly greater than mature A M . By contrast, there were no differences in the RSV Yield (neither in uptake of RSV nor in RSV progeny titer) in A M subpopulations from adult guinea pigs. Immature A M from juvenile guinea pigs had by far the highest RSV Yield of the A M subpopulations examined. DISCUSSION The working hypothesis for this study is that susceptibility of guinea pig A M to RSV infection depends on 2 intrinsic host factors, namely, the age of the host animal and the state of cell maturation. The main objective of this study was to determine and compare the in vitro RSV infection of A M subpopulations from adult and juvenile guinea pigs in terms of RSV Yield, defined as the RSV progeny per RSV-immunopositive cell. Distribution ofAM subpopulations Due to difficulties in obtaining clinical samples from the lower respiratory tract of infants and young children, examination and analyses of subpopulations of A M from children have not been investigated. To my knowledge, the distribution of A M subpopulations between different age groups or sexes (see Chapter 7) has not been investigated in animal models. The data obtained in this study showed that the distribution of mature, intermediate and immature A M between adult and juvenile guinea pigs is similar. In Chapter 5, based on the localization of RSV antigens in the lysosomes of mature A M as opposed to the cytoplasm in intermediate and immature A M , it was speculated that perhaps the propensity of children to severe RSV disease 61 may be explained by a proportionately greater distribution of A M towards the immature subpopulation. Although data from the studies of Ferro et al. (201) suggest that the majority of A M from the normal adult lung is composed mainly of the mature subpopulation, this phenomenon is not observed in the guinea pig. The observations in this animal model do not support the hypothesis that young children are more prone to severe RSV disease because of having a higher proportion of immature A M . As there is little or no published literature on the interaction of age of the host animal and A M subpopulations, it is presently unclear whether the data obtained are specific to guinea pigs or whether they can be extrapolated to other species. RSV Yield between Adult and Juvenile guinea pigs In this thesis, the term \"RSV Yield\" was introduced and defined as the ability of a cell to internalize RSV and its ability to support viral replication. By this definition, the measures of viral uptake (as determined by RSV immunostaining) and replication (as determined by plaque assay) in the A M are correlated and incorporated into one entity. Therefore RSV Yield is a reflection of viral replication within RSV-immunopositive A M based on immunostaining of RSV antigens. In this experiment, the RSV Yield was greatest in immature A M from juvenile animals, suggesting that this subpopulation of A M is particularly conducive to RSV propagation. The data obtained in this study are consistent with and extend the findings of previous investigators who documented increased viral replication in macrophages from young animals (83,108,113). Together with these previous findings by others, the data presented in this study demonstrate that as monocytes and the relatively immature A M differentiate and mature, they lose their tropism for RSV. Immature A M are more susceptible to RSV infection compared to the more mature cells in juvenile guinea pigs and also when compared to immature A M from adult animals, but this increased susceptibility does not appear to be related to differences in 62 distribution of A M subpopulations between the ages. Furthermore, these in vitro findings, when considered in the context of the in vivo observations by Hegele et al. (207), suggest that young animals may be highly susceptible to RSV infection of A M , with potential effects on more extensive viral replication, suppression of host lung defenses, and production of more severe lung injury. C H A P T E R 7 E F F E C T O F S E X O F H O S T A N I M A L 63 INTRODUCTION There are many risk factors known to predispose young children to RSV bronchiolitis. Children born prematurely, with underlying cardiac and/or pulmonary disease seem to be particularly susceptible to clinically severe RSV infections. In addition, the absence of breastfeeding, low levels of maternal antibodies, sex (male), crowding and being in daycare also increases the prevalence and severity of illness during the epidemic months (1). This study examines the effect of sex of the host animal on in vitro RSV infection in guinea pig A M . Epidemiological data of RSV-infected children indicates that male infants have a higher risk of RSV bronchiolitis compared to females with a ratio ranging from 1.4:1 to 1.9:1 (31-33). Furthermore, the more severe forms of RSV-bronchiolitis seen more frequently in male infants than female infants may be associated with eosinophilia during illness (209). The reasons and mechanisms responsible for these apparent sex-related disparities are not understood. Speculations include sex-related differences in fetal lung development that result in earlier maturation of the female lung than in males (210) and differences in pulmonary surfactant production between the sexes (211). The investigations in this study examined the issue of sex differences observed in RSV-bronchiolitis from 2 perspectives: the RSV Yield and the cytokine profile of infected A M . First, based on our earlier findings that immature A M are more susceptible to RSV infection than mature A M , the following questions were addressed: Do male guinea pigs have more immature A M than females? Do A M from male guinea pigs have a greater RSV Yield than females? Secondly, because a major role for cytokines has been implicated in RSV-induced inflammation (95, 116), the production of these pro-inflammatory cytokines by A M due to RSV infection was examined: T N F a which has anti-viral activity (212, 64 213), IL-6 which plays a major role in the initiation of the humoral arm of immunosurveillance, and IL-8, a major chemoattractant for neutrophils (69). HYPOTHESIS AND SPECIFIC AIMS Therefore the Working Hypothesis of this study is: A M from male guinea pigs are more susceptible to RSV infection and produce more inflammatory cytokines than A M from female guinea pigs. The Specific Aims of this study are: 1. To determine and compare the uptake of RSV in A M subpopulations between male and female guinea pigs. 2. To determine and compare viral replication of RSV in A M subpopulations between male and female guinea pigs. 3. To determine and compare RSV Yield in A M subpopulations between male and female guinea pigs. 4. To examine the secretion of IL-6, IL-8 and T N F a in subpopulations of A M from male and female guinea pigs. 65 EXPERIMENTAL PROTOCOLS Study Design The design of this study is shown in Figure 7.1. Based on the data obtained in the E M study and the age effect study, it was apparent that the intermediate and immature subpopulations of A M responded in a similar manner to RSV. Therefore, for this experiment, intermediate A M were isolated together with the immature cells by omitting the 20% metrizamide layer in the density gradient column. Juvenile female (n=4) and male (n=4) guinea pigs were sacrificed and heterogeneous A M were harvested by bronchoalveolar lavage. The heterogeneous A M were fractionated into mature and immature subpopulations and were exposed to RSV (m.o.i.=3). At 24 hours post-infection, the cell supernatants were used for cytokine assays and the cells were used for immunostaining for proportion of RSV-immunopositive A M and plaque assay for RSV progeny. 66 RSV infection (m.o.i.=3) * 1. Immunocytochemistry 2. Plaque Assay 3. IL-6, IL-8, TNFoc profile Figure 7.1 Experimental design for study of effect of gender of host animal on in vitro RSV infection in guinea pig A M . 67 Detection of IL-6 & IL-8 by ELISA A protocol established by Dr. T.R. Bai and associates (McDonald Research Laboratories) was utilized for the analyses of IL-6-like and IL-8-like proteins in supernatants from guinea pig A M . For IL-6 ELISA assay, monoclonal anti-human IL-6 antibody (4|ig/mL) was used as a capture antibody with biotinylated polyclonal anti-human IL-6 antibody (25ng/mL) as the detection antibody. Likewise, a monoclonal anti-human IL-8 antibody (4 u,g/mL) was used in conjunction with a biotinylated polyclonal anti-human IL-8 antibody (20 ng/mL) for the IL-8 ELISA assays. Recombinant human IL-6 and IL-8 were used in the respective assays for calculation of standard curves that were generated for each set of samples assayed. A l l antibodies and recombinant proteins for ELISA assays were obtained from R & D Systems. Plate Preparation High binding polyvinyl chloride plates (Corning Costar, Corning, N Y ) were coated with 100 |iL/well of the capture antibody overnight. Using a multi-channel pipettor, the plates were washed thrice with Wash Buffer8 to remove unadsorbed excess capture antibodies. 300 | i L of Blocking Solution9 were added to each well and incubated for one hour to minimize non-specific binding of cytokines to the plates. Again the plates were washed thrice using Wash Buffer, dried under vacuum and stored at 4\u00C2\u00B0C until used. Assay Procedure Dilutions of standards were carried out in polypropylene tubes. Samples of culture supernatant from guinea pig A M (2 x 2x106 A M per well per animal) and standards (100 |iL/well) were added and mixed gently by tapping the plate frame. The plates were sealed and 8 Wash Buffer: 0.05% Tween-20 (Sigma-Aldrich) in PBS, pH 7.4 9 Blocking Solution: 1% BSA (Sigma-Aldrich), 5% sucrose (Fisher Scientific), 0.05% NaN3 (Sigma-Aldrich) in PBS, pH 7.4 68 incubated for 2 hours. Washing, as described above, was carried out and the biotinylated detection antibody (100 |iL/well) was added and incubated for 2 hours. The aspiration/washing step was repeated and 100 |xL/well of streptavidin HRP (Zymed, South San Francisco, CA) was added and incubated for 20 minutes. A final aspiration/wash procedure was performed and 100 fiL/well of Substrate Solution 1 0 was added to each well and incubated for 20-30 minutes, protected from direct light. The reaction was stopped by the addition of 50 (ilVwell of Stop Solution1 1. The absorbance (A405) of each well was read within 30 minutes using a microplate reader set to a wavelength of 405 nm. Calculation of Results To calculate assay results, the zero standard absorbance was subtracted from the mean of the sample duplicate readings. A standard curve from each set of experiments was obtained by plotting the A405 against the concentrations of the standards. The equation of the best-fit line was determined and the concentration of the cytokine measured can be calculated based on its A405 reading and the best-fit equation. Statistical Analysis The data were presented as means + S E M of values. Data were analyzed using an A N O V A and the Bonferroni method was used for correcting multiple comparisons. A value of p<0.05 (two-tailed) was considered statistically significant. 1 0 Substrate Solution: 1:1 mixture of H 2 0 2 : Tetramethylbenzidine (Medix Biotech, San Carlos, CA) 1 1 Stop Solution: 1 M H 2 S0 4 (BDH Chemicals) 69 Detection of TNFa by Bioassay In our laboratory, Dr. R.R. Schellenberg and colleagues have used this protocol for the detection of T N F a released by A M from guinea pigs (214). The protocol, also known as L929 cell cytotoxicity assay, is a modified version of that described in Current Protocols. The proportion of lysis in the TNF-sensitive L929 cells indicates the level of bioactive TNF in the culture supernatant from guinea pig A M . To confirm the expression of bioactive TNF by guinea pig A M , a neutralization test was performed by using polyclonal rabbit anti-mouse T N F a neutralizing antibody (Genzyme, Cambridge, M A ) . Preparation of L929 fibroblasts A large batch of L929 fibroblasts was cultured in several T75 flasks until confluent. The cells were pooled together by trypsinization and washed in RPMI medium supplemented with 10% FBS. After washing by centrifugation (10-minute centrifugation at 400 x g, room temperature), the cell pellet was resuspended in supplemented RPMI. A cell count was performed and cell viability was evaluated by the trypan blue exclusion test. The cell density was adjusted to a final concentration of 4x10 s cells/mL. The fibroblast suspension was added to each well of a 96-well flat-bottom microtiter plate, at 100 |iL/well and incubated overnight at 37\u00C2\u00B0C in a 5% CO2 humidified incubator. In general, the L929 fibroblasts plated at this density yielded confluency the next day. Assay Procedure The confluency of the L929 cells in 96-well microtiter plates were checked the following day before proceeding with the assay as this is important for reproducible assays. The culture media was aspirated from each well of the microtiter plate using an 8-well aspirator with extra 70 care so as not to damage the monolayers with the aspirator tips. Recombinant human T N F a (rhTNFRa; R & D Systems) was used as a standard for calibration. Using an 8-channel pipettor, 2-fold serial dilutions of standards (ranging from 5 pg/mL to 5 ng/mL rhTNFRa) and culture supernatants from guinea pig A M (neat to 1:32 dilutions) were performed in the 96-well microtiter plates such that the final volume in each well contained 50 \iL. A l l standards and samples were plated in duplicate. One set of standards was prepared for every assay performed. The plates were incubated overnight (18 hours) in a 37\u00C2\u00B0C, 5% CO2 humidified incubator in the presence of the metabolic inhibitor actinomycin D (2.5 u.g per well, Sigma). Following overnight incubation, supernatant from each well was aspirated off and the L929 cells were washed with 200 uL of PBS. The wells were stained with 50 [iL of Crystal Violet solution1 2 for 10 minutes at room temperature. The wells were washed using cold tap water and excess water was removed by a sharp flicking of the plate. Care was taken so as not to hit the plates against any hard surfaces and thereby dislodging the adherent cells. The microtiter plates were inverted over absorbent paper and allowed to dry. Scoring for T N F a activity The Crystal Violet stains from the L929 fibroblasts were eluted by the addition of 100 U.L of 100% ethanol. The plates were read immediately with a microtiter plate reader at an absorbance of 595 nm. The mean absorbance reading of the negative control values was subtracted from each well. The resulting data of the standards were graphed by plotting the cellular response values (Y-axis) versus the T N F a concentrations. This calibration graph was used to determine the saturating maximal response value and the linear portion of the dose-response curve. The amount of T N F a required to stimulate a half-maximal response normally 1 2 Crystal Violet solution: 0.05% Crystal Violet in 20% Ethanol 71 falls within the linear portion of the dose-response curve and was used to define a unit of activity. The saturating maximal response value and linear portion of the dose-response curve for each set of serially diluted samples were obtained similarly by graphing the cellular response versus the reciprocal dilution. The half maximal response and the corresponding reciprocal dilution were calculated from the equation of the linear portion of the dose-response curve. The amount of T N F a present in the sample supernatant is equivalent to the product of its reciprocal dilution at half-maximal response and the corresponding amount of T N F a from the standard controls. Statistical Analysis The data were presented as means \u00C2\u00B1 S E M within the same animal in each experiment. Data were analyzed by A N O V A and the Bonferroni method was used for correcting multiple comparisons. A value of p<0.05 (two-tailed) was considered as statistically significant. 72 RESULTS Distribution of A M subpopulations between the Sexes The distribution of A M subpopulations from young male and female guinea pigs was examined and is presented in Table 7.1. There were no significant differences in the distribution of A M subpopulations between the sexes. Table 7.1 Distribution of A M subpopulations in juvenile male and female guinea pigs. 1 email' Male Mature A M 47.9 \u00C2\u00B14.3 47.3 \u00C2\u00B14.8 Immature A M 52.1 \u00C2\u00B14 .3 52.7 \u00C2\u00B14 .8 Data represent the proportion (mean \u00C2\u00B1 SEM) of mature and immature A M from young male (n=4) and female (n=4) guinea pigs. 73 Effect of Sex on RSV Immunopositivity The data describing the impact of sex on RSV infection are presented in Figure 7.2. In comparison to A M obtained from male guinea pigs, these results indicated that female guinea pigs had a higher proportion of RSV-immunopositive A M in both mature (female vs. male: 13.4 \u00C2\u00B1 0.6% vs. 9.9 + 0.4%, p<0.004) and immature (female vs. male: 25.1 + 0.8% vs. 21.2 \u00C2\u00B1 0.9%; p<0.04) subpopulations. Moreover, regardless of the sex of the host animal, the proportion of RSV-immunopositive cells within the immature subpopulation was significantly higher than those in the mature subpopulation (p<0.002). Effect of Sex on RSV replication Quantification of viable RSV progeny from RSV-exposed A M was determined using viral plaque assay. These data are presented in Figure 7.3. There were no significant differences in RSV progeny cultured in A M between male and female guinea pigs. Despite a significantly lower proportion of RSV-immunopositive cells from male guinea pigs, these cells contained similar amounts of RSV progeny as compared to the female cells. A significantly greater number of RSV progeny were isolated by viral plaque assay in immature A M compared to mature A M from both male and female guinea pigs (p<0.0003). Figure 7.2. Effect of sex of host animal on RSV immunopositivity. Data represent percentage of RSV immunopositive A M (mean \u00C2\u00B1 SEM) in mature (white bars) and immature (black bars) subpopulations from juvenile male and female guinea pigs, 'a' p<0.04 immature A M : female vs. male 'b' p<0.004 mature A M : female vs. male 'c ' p<0.002 mature vs. immature A M from both male and female guinea pigs. Figure 7.3 Effect of sex of host animal on RSV replication. Data represent RSV progeny per million A M exposed to RSV in mature (white bars) and immature (black bars) subpopulations from juvenile female and male guinea pigs, 'a ' p<0.002 mature vs. immature A M from both male and female guinea pigs. 75 Effect of Sex on RSV Yield The RSV Yield of A M between the sexes is presented in Table 7.2. These data indicate that immature A M were significantly more susceptible to RSV than mature cells. In the mature subpopulation, A M from male guinea pigs were significantly more susceptible to RSV than those from female guinea pigs (male vs. female: 41.3 \u00C2\u00B1 2.9 vs. 28.0 \u00C2\u00B1 2.8, p<0.009). By contrast, there were no significant differences between the sexes in the subpopulation of immature A M . Table 7.2 RSV Yield in A M from male and female guinea pigs. Female Male Mature A M 28.0 \u00C2\u00B12 .8 4 1 . 3 \u00C2\u00B1 2 . 9 b Immature A M 292.9 \u00C2\u00B1 11.7a 313.0 \u00C2\u00B132 .2 a Data represent the RSV yield (mean \u00C2\u00B1 SEM) of A M subpopulations from young guinea pigs. 'a ' p<0.002 mature A M vs. immature A M from both male and female animals, 'b' p<0.009 female vs. male in mature subpopulation of A M . Effect of Sex on Cytokine Production Cytokines released into cell supernatants following in vitro RSV infection were detected by ELISA (IL-6 and IL-8) and bioassay (TNFa). The levels of IL-6-like proteins, IL-8-like proteins and T N F a produced by subpopulations of A M from male and female guinea pigs are presented in Table 7.3. A l l 3 cytokines were upregulated in RSV-infected cells compared to sham-infected cells (see Appendix). However, there were no differences in the levels of cytokines released between A M subpopulations or between the sexes. 76 Table 7.3 Cytokine profile of guinea pig A M following RSV exposure. Cytokine A M Female Male IL-6 (pg/mL) Mature 349 \u00C2\u00B1 6 7 378 \u00C2\u00B1 6 8 Immature 331 \u00C2\u00B1 6 8 350 \u00C2\u00B1 8 2 IL-8 (pg/mL) Mature 627 \u00C2\u00B1160 730 \u00C2\u00B1 5 9 Immature 863 \u00C2\u00B1 206 871 \u00C2\u00B1108 T N F a (ng/mL) Mature 8.7 \u00C2\u00B12.1 8.5 \u00C2\u00B12.5 Immature 10.4 \u00C2\u00B13.7 13.1 \u00C2\u00B14.6 Data represent the amounts of IL-6, IL-8 and T N F a (mean \u00C2\u00B1 SEM) produced by RSV-exposed A M from male and female guinea pigs. Summary of Results The data obtained in this study are summarized as follows. The distribution of A M subpopulations in juvenile guinea pigs is similar between the sexes. Regardless of the sex of the host animal, immature A M are more susceptible to RSV infection and manifest a higher RSV Yield than mature A M . Although cells from female guinea pigs appear to be more susceptible to RSV infection than male guinea pigs, the quantity of viable RSV progeny isolated from these cells are similar between the sexes. Finally, examination of the cytokine profile revealed similar levels of RSV-induced IL-6, IL-8 or T N F a produced by different A M subpopulations from male and female guinea pigs. 77 DISCUSSION The working hypothesis in this study is that A M from male guinea pigs are more susceptible to RSV infection and produce more inflammatory cytokines than A M from female animals. This study was an attempt to determine if there was any disparity between the sexes of the guinea pig A M response to RSV, as a possible mechanism of the apparent propensity for boys to develop more clinically severe RSV infections than girls. This was achieved by examining the distribution of A M subpopulations between the sexes, the susceptibility of A M subpopulations to RSV (as reflected by RSV Yield) and the cytokine response of A M to RSV. Effect of sex on distribution ofAM subpopulations Analysis of the distribution of A M subpopulations from male versus female guinea pigs did not indicate any statistically significant differences. These findings do not support the possibility that males are more at risk for RSV-bronchiolitis because of an increased proportion of immature A M which are relatively more susceptible to in vitro RSV infection. There are no published data on similar analysis of A M subpopulations between the sexes in humans or other animal models. While epidemiological data suggest that young males are more at risk to RSV-bronchiolitis, such differences have not been reported in animal studies. Microscopic examination comparing histological changes in the lungs of RSV-infected male versus female guinea pigs may provide further information concerning the risk differences observed in male versus female infants. Regardless, the data in this study do not show any differences in the distribution of A M subpopulations between male and female guinea pigs. Effect of sex on RSV Yield in AM subpopulations While the immunostaining data indicated a higher proportion of RSV-immunopositive cells from female versus male guinea pigs, the plaque assay data showed that the quantities of 78 RSV progeny from both sexes were similar. Therefore, the RSV yield, defined as the quantity of RSV progeny per million A M that were exposed to RSV, was higher in the subpopulation of mature A M from young male guinea pigs compared to the respective cells from the female guinea pigs. Contrary to the initial speculation that male guinea pigs may have a greater uptake of RSV (based on expression of RSV antigens determined by immunostaining), these data showed that mature A M from male guinea pigs are capable of an enhanced support of RSV replication rather than increased viral uptake. These findings suggest that male guinea pigs may not be as efficient as female guinea pigs in restricting viral replication or inactivating RSV, presumably due to differences in fetal lung development between the sexes. It is conceivable that differences between male and female guinea pigs may involve cellular mechanisms of virus clearance. Whether a similar phenomenon occurs in humans remains to be elucidated. Effect of sex on RSV-induced cytokine production by AM subpopulations Cytokines play major roles in the modulation of inflammatory and immune responses. These molecules are often referred to as a \"two-edged sword\" because of their roles in protection as well as pathogenicity and the subsequent structural damage brought about by increased production. The relationship between RSV-induced cytokine production and the sex of the host animal has not been previously investigated. Examination of the cytokine response of juvenile male and female guinea pigs infected by RSV may aid in the understanding of increased severity of RSV-bronchiolitis seen in infant boys compared to infant girls. IL-6, IL-8 and T N F a were chosen to be analyzed in this study because of their major roles in the complex network of cell-cell communication, the activation of A M and previous reports describing RSV-induced increases in secretion of these cytokines by A M (114,116). In this study, the expression of these cytokines by guinea pig A M was upregulated in a similar manner when exposed to RSV; however, the results showed no differences in these cytokine levels between the sexes in guinea 79 pigs. This observation implies that the differential susceptibility of A M to RSV between the sexes has no effect on viral-induced production of these cytokines. A major limitation of the in vitro system used in this study is the lack of cell-cell interaction and the interaction between cytokines. Therefore, an experimental system that enables such interactions to be examined would be highly valuable in the understanding of the complex and dynamic network of cytokines involved over the course of an episode of RSV-bronchiolitis. Despite limitations, the data in this study show that the guinea pig is a useful animal model for studying human RSV infection. In conclusion, mature A M from male guinea pigs are more susceptible to RSV infection and support more viral replication compared to A M from female guinea pigs. However, these differences observed between the sexes were not associated with differences in distribution of A M subpopulations or differences in the cellular production of IL-6, IL-8 or T N F a induced by RSV infection. 80 C H A P T E R 8 R S V I N T E R A C T I O N W I T H PM10 INTRODUCTION Over the last 2 decades, there has been a remarkable increase in the prevalence of allergic respiratory diseases and it is hypothesized that environmental, and not genetic, factors are responsible (70). At the same time, for unexplained reasons, the prevalence of R S V bronchiolitis has increased 2.4-fold within 16 years (4). To obtain an improved understanding about the potential contribution of environmental factors in these marked increases in the prevalence of respiratory diseases, scientists have examined the role of air pollution, as one environmental factor. In some studies (215, 216), air pollution has been implicated as a possible protective factor in preventing development of atopy and asthma in later life. On the other hand, data from epidemiological studies indicated that air particulates caused adverse pulmonary health effects (217) as well as increased the susceptibility of children to RSV and other airway infections (218, 219). One plausible explanation for the increase in the prevalence of allergic respiratory diseases over the last twenty years is known as the \"Hygiene Hypothesis\". This term was coined by David P. Strachan (72) who proposed that improved household amenities and the higher standards of personal cleanliness have significantly reduced the incidence of various childhood infections. This in turn led to increased susceptibility to become sensitized to harmless allergens in later life. The basis for this hypothesis is that childhood viral or bacterial infections, which induce Thl type response and long-lasting immunological memory, could potentially prevent the Th2 immune responses in individuals susceptible to atopy and asthma. Hence common childhood infections are implicated as a protective factor against sensitization to allergens later in life. 81 Numerous reports appear to support the possibility that common childhood infections could play a protective role in development of allergic diseases. Results from various cohort studies (220-222) showed that young children who attend daycare or have older siblings are at a lower risk of atopic diseases. Other studies carried out in Guinea-Bissau (223) and in Scotland (224) provided epidemiological evidence that measles infection could reduce the risk of allergies, von Mutius and colleagues studied children from East and West Germany (considered a reasonably genetically homogeneous population) soon after the Reunification of 1989 and found that East German children had a higher frequency of respiratory disease but a lower prevalence of asthma compared to their West German peers (215). In a highly influential paper, Shirakawa et al. (225) showed that a positive tuberculin response in BCG-vaccinated schoolchildren in Japan correlated with a lower incidence of atopic conditions. Moreover, a similar inverse relationship was proposed between hepatitis virus infection and atopy (226). Taken together, the implication of these studies is that viral and bacterial infections during childhood might afford protection against allergic sensitization. However, there are others (227, 228) who dispute the apparent protective effects of childhood infections against atopic diseases. In contradiction to earlier studies (223, 224), data from Paunio et al. (227) showed that naturally acquired measles infection in Finnish children was associated with increased prevalence of atopic conditions. Furthermore, investigations on tuberculin response by Strannegard et al. (229) did not confirm the findings of Shirakawa and coworkers (225). Air pollution has been incriminated as an important health risk factor. For example, acute episodes of air pollution have been associated with increased mortality due to cardiovascular disease (230, 231). Epidemiological studies have reported that air particulates cause adverse cardiopulmonary effects especially in those with pre-existing lung diseases and are also associated with increased incidence of pneumonia and airway hyperreactivity (217). Other 82 studies have also suggested that pollution may influence the incidence of viral infections in children living in rural versus urban environment (218, 219). While the evidence for the pathogenic role of air pollution in lung disease is compelling, there have been numerous published studies that suggest a possible protective role for air pollution. The Reunification of Germany provided a unique opportunity to study the effects of air pollution on the development of allergic disorders in two genetically similar populations that had been exposed to different environmental conditions over four decades. Several studies of children and adults have demonstrated a significantly lower prevalence of bronchial hyperresponsiveness, hay fever and atopy in the more polluted East German cities compared to the relatively cleaner West German cities (215, 232, 233). Similar evidence was also obtained in studies performed in the Baltic region where schoolchildren living in the more polluted Poland demonstrated a lower prevalence of atopic sensitization and asthma compared to those in relatively less polluted western Sweden (234,235). The goal of the current study was to examine the interactive effects of RSV infection and air pollution (using PM10) on guinea pig A M functions. Recent studies by Becker et al. (193) showed that PM10 inhibits the ability of human A M to take up RSV. However, earlier studies demonstrated that RSV infection severely diminished the phagocytic ability of mouse A M (114). To understand how RSV and PM10 might interact and affect A M functions, we proposed that the A M response to RSV and PM10 is dependent on the sequence in which A M were exposed to these agents. The following specific questions were considered: Does PM10 exposure protect A M against RSV infection? Does R S V infection exacerbate a pre-existing PMlO-induced cytokine response of A M ? Does RSV protect against effects of acute PM10 exposure? Does exposure to PM10 exacerbate a pre-existing RSV-induced inflammation? Based on these considerations, in vitro experiments using guinea pig A M were designed to simulate 2 scenarios: RSV infection of individuals in high pollution areas (PM10+RSV) and RSV-infected individuals 83 exposed to an acute episode of air pollution (RSV+PM10). This was achieved by sequential exposure of A M to 2 environmental agents, R S V and PM10. In the former case, experiments were performed on A M that had been pre-treated with PM10. In the latter case, RSV-infected A M were subjected to a subsequent treatment of PM10. These 2 situations were analyzed in 3 aspects: (1) ability of A M to phagocytose PM10, (2) the R S V Yield based on ability of the A M to take up virus and support viral replication and (3) cytokine response of the A M to these stimuli. HYPOTHESIS AND SPECIFIC AIMS On the basis of the rationale presented above, the Working Hypothesis of this study is: The outcome of the interaction of R S V and environmental particulates in A M is dependent on the sequence in which the cells were exposed to these agents. The Specific Aims of this study are: 1. To compare the effect of sequential PM10 and R S V exposure, and vice versa, on the phagocytic function of A M using flow cytometric analysis. 2. To compare the effect of sequential PM10 and R S V exposure, and vice versa, on R S V uptake by A M . 3. To compare the effect of sequential PM10 and R S V exposure, and vice versa, on R S V replication. 4. To compare the effect of sequential PM10 and R S V exposure, and vice versa, on R S V Yield. 5. To measure cytokine production (IL-6, IL-8 and TNFa) of A M in response to sequential PM10 and R S V exposure, and vice versa. 84 EXPERIMENTAL PROTOCOLS Study Design The design of this study is shown in Figure 8.1. In this study, heterogeneous A M were obtained by bronchoalveolar lavage from female juvenile guinea pigs (n=4). Data from the previous study (Chapter 7) has indicated that neither the gender of the host animal nor the maturation stage of the A M influenced the cytokine profile of RSV-infected A M . Therefore in this experiment, A M were used and examined as a heterogeneous population. * \u00E2\u0080\u00A2Scatter Profile by Flow Cytometry \u00E2\u0080\u00A2RSV labeling by Flow Cytometry \u00E2\u0080\u00A2Plaque Assay \u00E2\u0080\u00A2IL-6, IL-8, TNFa profile Figure 8.1 Experimental design for study of PM10-RSV interactions in guinea pig A M . 85 Heterogeneous A M from each guinea pig were subjected to five different treatments: exposure to PM10 followed by RSV infection (PM10+RSV), RSV infection followed by PM10 exposure (RSV+PM10), exposure to PM10 only (PM10), exposure to RSV only (RSV) and the negative control (i.e., exposure to neither PM10 nor RSV) (NEG). At 24 hours post-treatment, the A M were used for plaque assay to quantify RSV progeny and flow cytometric analysis of proportion of RSV-infected cells and cell scatter properties while the cell supernatants were used for ELISA (IL-6 and IL-8 detection) and bioassay (TNFa). Several reports have demonstrated that the cytokine response of A M to environmental particulates might be due to the presence of endotoxin on particles (193, 236). Results from preliminary experiments indicated trace amounts of endotoxin (0.05 EU/100 | ig PM10) in the batch of particulates used in this study and this amount of endotoxin is unable to provoke a cytokine response in guinea pig A M by using a similar method of in vitro exposure (237). In addition, the dose response and viability of A M subpopulations to various concentrations of PM10 was examined. These results showed a minimal toxic effect on the A M (>90% cell viability by trypan blue exclusion test) and a maximal cytokine response with 100 ug/mL PM10 (238). Consequently, a PM10 concentration of 100 |J.g/mL was selected for use in the present study. Furthermore, the association of PM10 with A M was examined using both light and electron microscopes. Light microscopic examination of cytospin preparations (method is as described in Chapter 6) of A M exposed to PM10 localized the particles within the cytoplasm (Figure 8.2). Using techniques as described in Chapter 5, PMlO-exposed A M were embedded in Epon and examination under transmission electron microscope revealed PM10 within lysosomal compartments of A M (Figure 8.3). Figure 8.2 PMlO-exposed A M in hematoxylin-eosin stain. Particulates (arrow) are located within cytoplasmic regions of A M . Bar represents 25 urn Figure 8.3 Electron micrograph of an A M containing PM10 (arrow) within a lysosome. Bar represents 100 nm. 87 Particulate Matter (PM10) The PM10 used in this thesis work was obtained from Environmental Health Canada, Ottawa (EHC-93, kindly provided by Dr. R. Vincent of Health Canada). Airborne urban particles were collected from the outdoors using a single-pass air-purificator by vacuum and were sieved through a 36-(xm mesh filter before using for experimental purposes. The median diameter of this preparation of fine particles is 0.35 (xm but also retained a coarse component (approximately 7 to 15 Jim). The chemical components of EHC-93 consisted mainly of polycylic aromatic hydrocarbons, ions and metals and are listed in Table 8.1 (239). The individual components of this complex mixture of organic and inorganic compounds responsible for adverse health effects have not been identified. However, the soluble metal components in ambient PM10 have been proposed as the major contributor in stimulating the production of cytokines by A M (240). In addition, it has been suggested that the PMlO-associated metals with redox potential (for example, iron, vanadium or copper) play significant roles in contributing to the toxicity of particulate matter (241-243). 88 Table 8.1 Chemical components of EHC-93 Extracted and modified from Vincent et al. (239), with permission. Polycyclic Aromatic Hydrocarbons Anthracene Benzo(a)anthracene Benzo(b)fluoranthene Benzo(ghi)perylene Benzo(a)pyrene Fluoran there Phenanthrene Pyrene Ions Hydrogen ion Sulfate ion Metals Aluminum Chromium Copper Iron Lead Magnesium Nickel Vanadium Zinc Hg/g particle 0.54 1.10 2.78 1.52 0.95 2.47 1.83 2.11 jlg/g particle 0 45 x 103 u.g/g particle (% solubility in water) 10 x 103(2%) 42 (3%) 845 (17%) 15 x 103 (1%) 7 x l 0 3 (4%) 7 x 103(14%) 67(7%) 90(0%) 10 x 103(46%) PM10-RSV Exposure The EHC-93 particles were prepared fresh as 100 |ig/mL particulate suspensions using RPMI media supplemented with 2% FBS and gentamycin. To minimize aggregates of particles, the PM10 suspensions were sonicated using a probe sonicator (VibraCell\u00E2\u0084\u00A2, Sonics & Materials, Danbury, CT) 3 times at 5 minutes intervals prior to use. 500 |J,L of PM10 particulate suspensions were added to 6-well plates containing approximately 2 x 106 adherent A M per well. The cells were exposed to PM10 particulates for 60 minutes in a 5% CO2 incubator at 37\u00C2\u00B0C. Following exposure, the excess PM10 suspensions were removed by washing twice with PBS and 1 mL of fresh medium was added to wells with no RSV infection. In wells where A M were subjected to both agents, the cells were treated with the first agent (PM10 or RSV) and washed 89 twice with PBS before being treated with the second agent (RSV or PM10) immediately following the first treatment. Each treatment was carried out for 60 minutes, with intermittent agitation to ensure even exposure, in a 37\u00C2\u00B0C 5% CO2 incubator. After removal of the second agent, 1 mL of fresh medium was added to each well and the cells were allowed to incubate in a 37\u00C2\u00B0C 5% C 0 2 incubator overnight. Flow Cytometry Flow cytometry was used to analyze the cells because of the ability of this technique to measure multiple parameters in large number of cells simultaneously. In this thesis, flow cytometry was utilized for the measurement of cell parameters (such as cell size and granularity) and the expression of RSV antigens defined by fluorescent antibodies. The samples were analyzed using the FACScan flow cytometer (UBC Biomedical Research Center, Flow Cytometry Facility). The term \" F A C S \" is Becton-Dickinson's registered trademark and is an acronym for \"Fluorescence-Activated Cell Sorter\". The FACScan uses an air-cooled argon gas laser with a fixed wavelength emission of 488 nm. It has fluorescence detectors that detect green, yellow-orange and red light. For the purpose of this thesis, fluorescein was used extensively for the green channel. The FACScan analyzes cells at the rate of several hundreds per second and 5,000 to 10,000 cells were acquired per sample. The samples were in mono-disperse suspensions whereby the cells passed single-file through a laser beam by continuous flow of a fine stream of the suspension. The measurement of the side scatter intensity, which is proportional to the granularity of the cell and reflection of particle ingestion, was obtained simultaneously with the fluorescence intensity of the fluorescent-labeled RSV-infected cells. The photomultiplier tube (PMT) voltage and compensation were set using cell surface staining controls (i.e., IgG of equivalent protein concentration) and the same quadrant markers were used for all experiments to facilitate inter-90 experiment comparisons. Data were saved to zip disks and analyzed with WinMDI (version 2.8), a graphics software. WinMDI is a freeware obtained from the World Wide Web (http://facs.scripps.edu). Using WinMDI, histograms and bivariate dot plots were generated upon data reanalysis to display the mean fluorescence intensity (MFI) frequency of A M with expression of surface and intracellular RSV antigens. Immunofiuorescent staining of RSV antigens The protocol described below was established empirically and optimized for combined surface and intracellular labeling of RSV antigens in guinea pig A M . Following treatment, the A M were scraped off 6-well plates using disposable sterile cell scrapers (Fisher Scientific). The wells were rinsed with 1 mL PBS and collected into the respective tubes. Since the anti-RSV antibodies (NCL-RSV3-FITC, Novocastra, UK) used in this protocol is a pool of monoclonal antibodies that detect both surface and intracellular antigens, a two-step immunostaining procedure was devised. The cells were washed in PBS and underwent centrifugation for 8 minutes at 1000 x g. A cell pellet was obtained by rapid decanting and was resuspended in 100 HL PBS. 10 u.L of FITC conjugated anti-RSV antibody (NCL-RSV3-FITC) was added to the cells and incubated for 30 minutes. Excess antibodies were washed off and the cells were fixed in 4% paraformaldehyde (Ted Pella) in PBS for 5 minutes. After washing, the cells were incubated in 0.5% Triton-X 100 (Sigma) for 10 minutes. Intracellular RSV antigens were labeled by a second incubation with the NCL-RSV3-FITC antibody for 30 minutes. The cells were washed in PBS and 0.2 mL crystal violet (2mg/mL PBS, B D H Chemical) was added to quench auto-fluorescence (244). After 5 minutes, a final wash with PBS was carried out and the cells were resuspended in 500 | i L PBS and stored on ice. A l l incubations were performed at room temperature unless otherwise stated. Flow cytometry was performed within 2 hours at the U B C Biomedical Research Center, Flow Cytometry Facility. The proportion of RSV-91 immunopositive cells was determined by the amount of green fluorescence emitted by the FITC conjugated anti-RSV antibody. Statistical Analyses Data were expressed as the mean value \u00C2\u00B1 SD. A N O V A analysis was performed with a Bonferroni correction for multiple comparisons. A value of p<0.05 (two-tailed) was considered as statistically significant. RESULTS The effects of sequential exposure of A M to PM10 and/or RSV were examined in this study. The objectives were to examine the impact of sequential PM10-RSV interaction with respect to the phagocytic ability of A M , ability of A M to take up R S V and support viral replication as well as the cytokine response of these cells. Effect of PM10-RSV interaction on the phagocytic ability of A M The ability of A M to phagocytose PM10 was determined indirectly by the measurement of side scatter that is proportional to the granularity of the cell and a reflection of particle ingestion. The effect of PM10-RSV interaction on A M granularity is summarized and presented in Figure 8.2. A M that were exposed to PM10 showed a significant increase in mean side scatter in comparison to the negative control A M (p<0.05) and in comparison to RSV-infected A M (p<0.04). There were no significant differences in mean side scatter between negative control A M and RSV-infected A M . In addition, there were no significant differences in mean side scatter between A M that were exposed only to PM10 and A M that were exposed to both agents. PM10+RSV RSV+PMIO PMIO RSV N E G Figure 8.4. Effect of PMIO-RSV interaction on A M granularity. Data represent side scatter (mean \u00C2\u00B1 SD) of guinea pig A M subjected to different treatments, 'a' p<0.05 N E G vs. PM10+RSV, N E G vs. RSV+PMIO, N E G vs. PMIO 'b' p<0.04 RSV vs. PM10+RSV, RSV vs. RSV+PMIO, RSV vs. PMIO 93 Effect of PMIO-RSV interaction on R S V Immunopositivity Typical histograms of RSV labeling by flow cytometry on HEp-2 cells and guinea pig A M with and without RSV infection are presented in Figure 8.5. The mean fluorescence of uninfected HEp-2 cells (Top Left) and guinea pig A M (Top Right) labeled with anti-RSV antibodies and IgGl (equivalent isotype control) are very similar. The population of RSV-immunopositive HEp-2 cells (>85%) is indicated by a right shift increase in mean fluorescence intensity (MFI, Bottom Left, anti-RSV vs. IgGl: 53.2 vs. 4.3). The spread of this cell population indicates that some HEp-2 cells are more heavily infected than others. Bottom right panel in Figure 8.3 shows a right shift of a minor proportion (>25%) of RSV-immunopositive guinea pig A M when labeled with anti-RSV antibodies (MFI anti-RSV vs. IgGl: 17.9 vs. 1.6). The data describing the effect of PMIO-RSV interaction on RSV infection are summarized and presented in Figure 8.6. A l l four treated groups showed a significantly greater proportion of RSV-immunopositive cells compared to negative control A M (p<0.007). The low proportion (1.98% \u00C2\u00B1 0.37% (mean \u00C2\u00B1 SD)) of RSV-immunopositive cells in the PMIO group may be attributed to non-specific staining. In comparison to A M in the RSV group, A M that were initially exposed to RSV followed by exposure to PMIO (i.e., RSV+PMIO group) showed a similar proportion of RSV-immunopositive A M . By contrast, A M that were initially exposed to PMIO followed by exposure to RSV (i.e. PM10+RSV group) showed a significantly smaller proportion of RSV-immunopositive A M (PM10+RSV vs. RSV: (4.5% \u00C2\u00B1 0.84%) vs. (10.5% \u00C2\u00B1 5.11%), p<0.05). In addition, for cells that were subjected to both treatments, the RSV-immunopositivity of A M is influenced by the sequence of exposure to these agents. The data indicated that RSV uptake by A M is suppressed if these cells had been exposed to PMIO prior to RSV (PM10+RSV vs. RSV+PMIO: (4.5% + 0.8%) vs. (12.6% \u00C2\u00B1 0.9%), p<2 x 10\"5). 94 HEp2 cells Anti-RSV fc Guinea pig A M RSV IgGl 10' FL-1 10J RSV-infected HEp2 cells IgGl -Anti-RSV 10* FL-1 10\u00C2\u00BB 10* RSV-infected guinea pig A M 10'* 10\" Figure 8.5 Representative histograms of HEp-2 cells and guinea pig A M with and without RSV infection. Top Left: HEp-2 cells labeled with anti-RSV antibodies (Red) and IgGl (Black line, no fill). Bottom Left: RSV-exposed HEp-2 cells labeled with anti-RSV antibodies (>85% RSV immunopositive cells, MFI=53.2) and control IgGl (MFI=4.3). Top Right: Guinea pig A M labeled with anti-RSV antibodies and IgGl. Bottom Right: RSV-exposed guinea pig A M labeled with anti-RSV antibodies (>25% RSV immunopositive cells, MFI=17.9) and control IgGl (MFI=1.6). Figure 8.6 Effect of PMIO-RSV interactions on RSV immunopositivity. Data represent percentage of RSV-immunopositive A M (mean \u00C2\u00B1 SD) as determined by flow cytometric measurement of FITC-linked RSV antibodies. 'a ' p<0.007 N E G vs. all other groups 'b' p<0.05 PM10+RSV vs. RSV V p<2xlO'5 PM10+RSV vs. RSV+PMIO 95 Effect of PMIO-RSV interaction on R S V replication The quantity of RSV progeny isolated from A M , as determined by plaque assay, is shown in Figure 8.7. Negative control A M and those that were exposed to PMIO alone did not propagate RSV progeny. A M that were exposed to both agents produced 3- to 9-fold less RSV progeny compared to A M that were exposed to RSV alone (PM10+RSV vs. RSV (mean \u00C2\u00B1 SD): (390.6 \u00C2\u00B1 205) vs. (3534 \u00C2\u00B1 1457), p<0.005; RSV+PMIO vs. RSV: (1009.4 \u00C2\u00B1 449) vs. (3534 \u00C2\u00B1 1457), p<0.02). However, the quantity of RSV progeny was not significantly affected by the sequence of exposure to RSV and PMIO. 6000 - i 5 5000 -^ 4000 -o = 3000 s 33 2000 1000 PM10+RSV RSV+PM10 PM10 RSV NEG Figure 8.7 Effect of PM10-RSV interaction on RSV replication. Data represent quantity of RSV progeny scored per million A M (mean \u00C2\u00B1 SD). 'a' p<0.005 PM10+RSV vs. RSV 'b'p<0.02 RSV+PM10 vs. RSV 96 Effect of PMIO-RSV interaction on R S V Yield RSV Yield, as explained in earlier studies, is a measurement of the amount of replicating virus per RSV-immunopositive A M as determined by immunolabeling of RSV antigens. RSV Yield can therefore be considered as a reflection of the robustness of viral replication within an infected cell. These data are summarized and presented in Table 8.2. A M that were exposed to RSV alone produced the highest RSV Yield (p<0.04). When A M were exposed to both RSV and PMIO, a 5-fold decrease in RSV Yield was observed. This reduction in RSV Yield was independent of the sequence by which A M were exposed to both RSV and PMIO. Table 8.2 Effect of PMIO-RSV interaction on RSV Yield. Is' Exposure 2'\"1 Exposure RSV \ ield RSV - 463 \u00C2\u00B1 346a PMIO RSV 88 \u00C2\u00B1 4 0 RSV PMIO 79 \u00C2\u00B1 37 Data represent RSV Yield (mean \u00C2\u00B1 SD) of A M in different treatments, 'a ' p<0.04 PM10+RSV vs. RSV, RSV+PMIO vs. RSV 97 Effect of P M 1 0 - R S V interaction on IL-6 production IL-6-like proteins released into cell supernatants following various treatments were detected by ELISA. The levels of IL-6-like proteins produced by A M are summarized and presented in Figure 8.8. Of the different treatments, A M that were exposed to RSV alone produced the most IL-6-like proteins (p<0.0002). Negative control A M and PMlO-exposed A M released similar amounts of IL-6-like proteins. Exposure of A M to PMIO significantly suppressed RSV-induced IL-6 production (PM10+RSV vs. RSV (mean \u00C2\u00B1 SD): (69.3 \u00C2\u00B1 6.4) vs. (327.9 \u00C2\u00B1 31.7), p<4 x 10 - 6; RSV+PM10 vs. RSV: (11.3 \u00C2\u00B1 2.2) vs. (327.9 \u00C2\u00B1 31.7), p50% <30% 30-50% >50% Lyn ERK1-B1 Cdk2 \u00E2\u0080\u00A2 E R K 1 - A 2 ERK2-B1 Ck4 PKCC GRK2 Cdk7 ERK2-A3 Ck2a p38a Ck2a' Fyn-U \u00E2\u0080\u00A2 E R K 2 - A 2 p90 S6K Fyn-L p43 F A K Hpk MEK1 \u00E2\u0080\u00A2 J A K 1 p70 S6K L * M E K 6 P K B a p52 S6K Rsk2 P A K a P K C p l RAF1 Rskl ZIPK + Data confirmed by a different vendor using similarly treated cells. 119 DISCUSSION The working hypothesis for this study is that RSV infection of guinea pig A M affects intracellular protein kinases involved in molecular mechanisms of virus-cell interactions. The objectives of this study were to determine which kinases are expressed in guinea pig A M and to determine changes in the expression of the protein in response to in vitro RSV infection by a \"blast\" screen method. As a first step, these studies may help to delineate potential pathway(s) involved in the activation of these cells as a consequence of the viral infection. The data in this \"blast\" screen were obtained for one time point, at 18 hours post-infection, when late signals and/or signals that were activated and sustained over long periods could be detected. Therefore early and transient events were not examined in this analysis. Expressions of protein kinases were detected in RSV-infected A M and the control cell line, HEp-2 cells. Protein kinases in RSV-infected HEp-2 cells at 18 hours post-infection were predominantly downregulated. Trypan blue exclusion test on these cells indicated viability of over 95%, suggesting that while most of the cells are viable, 18 hours post-infection is not a suitable time point to detect upregulation of protein kinases induced by RSV infection. Since activation of protein kinases could occur very fast (seconds to minutes) following receptor-ligand binding, future studies of protein kinases in RSV infected HEp-2 cells should be carried out at earlier time points. In addition, it would be useful to test for the activities of kinases that show changes in expression. Protein kinases with phosphorylation activities at earlier times may act as key regulators in signaling cascades. Nonetheless, in this study, the HEp-2 cells indicated that the K P K S 1.0 is a feasible method for detecting expression of protein kinases in RSV-infected guinea pig A M . Both upregulation as well as downregulation of protein kinases were detected in RSV-infected A M . This suggests that the kinetics of R S V infection in guinea pig A M versus HEp-2 cells is quite different. From the data summarized in Table 9.2, several important pathways and 120 cross-talk between these pathways could be indicated. In this discussion, I would like to focus on 2 major pathways: (1) Mitogen-activated protein kinase (MAPK), and (2) Protein kinase B (PKB/Akt). (1) The MAPK pathway The data in this study clearly indicated RSV-induced changes in components of the M A P K pathway. The amount of M E K increased by 3.5 fold and that of ERK2 increased by approximately 3.0 fold following RSV infection of A M . These data are interesting in light of the studies by Chen et al. (290), who demonstrated activation of M E K in RSV-infected A549 cells. In addition, these data confirmed the upregulation of M E K as demonstrated by Monick et al. (295) and emphasized the significance of the M E K - E R K M A P K pathway as proposed by them. However, in this \"blast\" screen, the two upstream components in the M E K - E R K pathway - P K C and Raf, that were activated by RSV infection (295) appear to be downregulated. This discrepancy could be due to an intended \"feedback\" mechanism that downregulates the upstream elements of the pathway. Alternatively, it may be due to a time factor in the design of the experiment and future studies with early time points might ascertain temporal consequences. In macrophages, T N F a plays an important role in initiating activation of the M A P K cascade by binding to the p55 TNF receptor (CD120) (318). Although a specific RSV response element has been identified in the IL-8 promoter (279), it is unclear whether M A P K expression is mediated directly by RSV or i f a synergistic role of TNF-induced M A P K activation is implicated. Furthermore, whether the RSV G protein, which has structural homology to the TNF receptor (20), plays a role in this pathway remains to be elucidated. In addition, the data in this study indicated that levels of Rsk, a substrate for E R K activation also increased. RSV-mediated activation of Rsk has not been reported (as of February 2002), an increase in the level of Rsk expression suggests that the potential involvement of this 121 kinase should be considered. Rsk belongs to a family of 90 kDa ribosomal S6 kinases and being a substrate of E R K , mediates M A P K signal transduction. The role of Rsk has been implicated in proliferation, differentiation and cell survival (319, 320). More specifically, and interestingly, Rsk is involved in regulation of gene expression via association and phosphorylation of numerous transcriptional regulators including NFKB/IKBOC (321, 322). Shouten et al. demonstrated that IKBOC is a target for Rsk and phosphorylation of Rsk is essential for subsequent degradation of IKBOC (321). The role and the mechanisms of Rsk in regulation of RSV-induced expression of IL-6 and IL-8 via the M A P K pathway and activation of N F K B require further investigation. RSV interaction with host cells has been described as biphasic: viral binding which induces a transient activation of E R K through PKC\u00C2\u00A3 activation of M E K and viral replication which causes a more sustained activation of E R K through Raf activation of M E K (280, 295). Biphasic activation of host cell is not restricted to RSV. A similar phenomenon has been observed in Coxsackievirus B3 infection of HeLa cells; Luo et al. (323) demonstrated that the inhibition of M E K and E R K resulted in suppression of viral protein and replication as well as reduced cell apoptosis. As the protein lysates were prepared 18 hours post-infection, it is conceivable that the increased expressions of M E K , E R K and Rsk may be associated with late activation of A M by RSV. Future studies to inhibit RSV progeny release and determination of viral gene/protein involved in the facilitation of sustained activation of intracellular signal transduction will enhance the understanding of R S V - A M interaction and may contribute to therapies in RSV disease. 122 (2) The PKB pathway In this study, RSV infection of A M induced increased expression of the protein kinases PKBoc and p70 S6K. The activities of both these kinases were not detectable in control cells. So far neither the roles of PKBoc and p70 S6K nor any signaling pathways involving these 2 protein kinases have been implicated in RSV-cell interaction. PKBoc, a multifunctional serine/threonine kinase (324, 325), plays a key role in insulin metabolism (326, 327), in cell survival mechanisms (328, 329) as well as in various forms of human cancers (330-333). Activation of the phosphatidyl 3-OH kinase (PI3K) has been utilized as a survival strategy by various forms of cancer cells; consequently this survival signal pathway is a major target for the development of inhibitors used in drug interventions (334, 335). P K B was identified as the cellular homologue of the viral oncogene v-Akt from A K T 8 , a retrovirus found in rodent T-cell lymphoma (336). P K B is activated by various growth and survival factors and is a target of PI3K (337, 338). Following stimulation of growth factor receptors, PI3K is activated and generates phosphorylated phosphatidyl inositides (PI-3,4-P2 and PI-3,4,5-P3). These phosphoinositides bind with high affinity to the amino-terminal pleckstrin homology (PH) domain of P K B and recruits P K B to the plasma membrane (339). To fully activate P K B , subsequent phosphorylation of Thr308 in the activation loop and Ser473 in the carboxyl terminal hydrophobic site are required. The conformational change, as the result of the binding between the PH domain of P K B and PI-3,4-P2/PI-3,4,5-P3, unmasks the activation loop site and allows phosphoinositide-dependent kinase (PDK-1) phosphorylation at Thr308 (340, 341). While phosphorylation at Ser473 is essential for activation of P K B , despite extensive biochemical analyses, the upstream kinase for this hydrophobic site remains to be identified (342). Cellular substrates downstream of P K B have been identified. These include the metabolic regulators glycogen synthase kinase 3 (GSK3) (326) and 6-phosphofructo 2-kinase 123 (PFK-2) (343), the Bcl-2 related apoptosis regulatory protein B A D (344, 345) as well as the ribosomal protein S6 kinase (p70 S6K) (338). These identified P K B substrates contain the sequence motif RxRyz(S/T)(hy) where x is any amino acid, y and z are small residues other than glycine and hy is a hydrophobic residue (346). The consequences of these activated P K B substrates include control of glycogen metabolism and suppression of apoptosis (347). As alluded to earlier (see Introduction), apoptosis is a contentious issue in RSV research. The potential role of P K B in preventing apoptosis may provide insights to regulatory role of RSV in cellular apoptosis. Activated P K B promotes cell survival through various distinct pathways. P K B inhibits apoptosis by phosphorylating the pro-apoptotic Bcl-2 family member B A D . Uncoupling of B A D from the B A D / B C I - X L complex inhibits the apoptotic pathway. In addition, P K B can also suppress apoptosis via several other mechanisms such as phosphorylation of Caspase-9 (348), CREB (349), GSK -30 (350), the Forkhead transcription factor F K H R L 1 (351) and IKB kinase (352). Although it has been suggested that RSV-mediated apoptosis is independent of N F K B (280) and unrelated to Caspase-9 (307), whether P K B can promote cell survival in RSV-infected cells via any of the other aforementioned mechanisms remains to be tested. Furthermore, whether a RSV protein can directly activate P K B via PI3K as seen in other viruses (353, 354) has yet to be determined. Therefore, examining the phosphorylation activity of P K B , particularly at the residues Thr308 and Ser473, may provide useful information. The protein kinase p70 S6K deserves some attention in this study. In vivo, p70 S6K phosphorylates ribosomal protein S6 which plays a major role in the translation of mRNA transcripts encoding ribosomal proteins and protein synthesis elongation factors (355). p70 S6K is shown to be phosphorylated by P D K in a manner similar to P D K phosphorylation of PKB (356). Burgering et al. have demonstrated that p70 S6K is a target for P K B ; however, other protein kinase(s) upstream of p70 S6K and downstream of P K B may be involved (338). More interestingly, p70 S6K has been found in complexes with protein phosphatase 2A (PP2A) (357). 124 Recent studies have indicated that the RSV P protein may play an important role in the persistent activation of N F K B via its association with PP2A (280). Taken together, these data suggest that the role of the RSV P protein may have a more extensive role in virus-cell interaction than previously thought. Summary This \"blast\" screen has identified numerous protein kinases and potential signaling pathways that may be pertinent to the intracellular signal transduction activities during RSV infection of guinea pig A M . For the purpose of this thesis, the discussion focused on 2 major and distinct pathways ( M A P K and PKB) and 3 protein kinases (Rsk, P K B and p70 S6K) whose potential involvements in RSV-mediated signal transduction are novel. The significance of the M E K - E R K M A P K pathway in RSV-infected cells was reiterated and a potential role for Rsk in this pathway and the regulation of N F K B activation is implicated. It is conceivable that the PKB-p70 S6K pathway may provide further insights to the role of RSV in cellular apoptosis. Through the analysis of expressions of protein kinases, this discussion has highlighted the roles of a limited number of protein kinases; that of many other potential candidates (for instance PDK, F A K , CK2 etc.) have not been reviewed. It should be emphasized here that this study utilized the KPKS-1.0 which measured the relative abundance of protein kinases present in a sample. Although some of the protein kinases in SDS-PAGE gels indicated mobility shifts (changes in mobility due to phosphorylation are often correlated with changes in activity states), KPKS-1.0 does not measure kinase activity. Future experiments designed to examine the phosphorylation activities of these candidate protein kinases could improve our understanding of R S V - A M interaction. The availability of the completely sequenced RSV genome, comprehensive databases of signaling information, peptide libraries, protein-RNA, protein-DNA and protein-protein interaction systems, suggest a bright future for RSV research. 125 C H A P T E R 10 C O N C L U D I N G R E M A R K S RSV infections represent a major source of morbidity in young children. Despite 46 years of research since its discovery, the numerous issues of controversies concerning the mechanisms of RSV-induced disease remain uncertain. Moreover, recent trends of the increasing incidence of RSV bronchiolitis are alarming. This thesis examines the interaction of RSV and guinea pig A M in vitro, where the conditions can be well controlled. The results confirm that RSV causes a productive infection in A M . On the basis of these experiments, several major findings were obtained and are summarized below. 1. Preferential RSV replication depends on the stage of cell maturation and age and sex of the host animal in our guinea pig model. These results reflect a similar phenomenon observed in humans, demonstrate the relevance of the guinea pig model and provide new insights at a cellular level. Further E M studies are needed to provide quantitative information on the abundance of lysosomes in A M subpopulations from animals of different age and sex groups. Additional studies examining the mechanisms by which RSV might evade host immunity will yield relevant information on persistence of RSV infection in guinea pig lung and probable persistence in humans. 2. The interaction of PMIO and RSV in guinea pig A M resulted in suppression of RSV replication, and RSV-induced cytokine production. The mechanisms involved in the uptake of PMIO resulting in the aforementioned A M response to RSV are unclear. Future studies to determine the potential antiviral role of NO, for example, may provide insights to the mechanisms involved in the interaction between PMIO and RSV. In addition, in vivo studies of sequential exposure of these agents will be valuable in 126 determining the significance of reduction of viral replication and RSV-induced cytokine expression in the complexity of the host organism and their effects on virus clearance mechanisms and lung pathology. Cell signaling studies may also provide valuable information on the activation of A M by RSV-PM10 interaction. 3. The M E K - E R K M A P K and P K B pathways, as indicated by expression of the corresponding protein kinase members play crucial roles in the response of A M to RSV infection. Future studies on activation of protein kinases in these pathways (in particular Rsk, P K B and p70 S6K) are needed to ascertain their significance in RSV-mediated signaling in A M and elucidate the mechanisms involved in RSV-induced cytokine production and the controversies of apoptosis in RSV-infected cells. In conclusion, the studies in this thesis demonstrated that the A M response to RSV infection is dependent on intrinsic properties of the cell (i.e. maturation stage), host factors (i.e., age and sex) and environmental factors (for example, air pollution). The findings in this thesis, which form a basis for further studies on mechanisms of susceptibility to RSV infection, have provided new insights on R S V - A M interaction and the potential effects of these interactions on the pathogenesis of RSV disease in young children. 127 BIBLIOGRAPHY 1. Wohl M E B . Bronchiolitis in children. In: Epler GR, editor. Diseases of the Bronchioles. New York: Raven Press Ltd; 1994. p. 397-407. 2. Falsey AR, Walsh EE. Respiratory syncytial virus infection in adults. Clin Microbiol Rev 2000; 13(3):371-84. 3. Institute of Medicine. Prospect for immunizing against respiratory syncytial virus. In: Institute of Medicine. New vaccine development: establishing priorities. Washington, DC: National Academy Press; 1986. p. 299-307. 4. Shay D K , Holman RC, Newman RD, Liu L L , Stout JW, Anderson LJ . Bronchiolitis-associated hospitalizations among US children, 1980-1996. J A M A 1999; 282(15):1440-6. 5. Boyce TG, Mellen B G , Mitchel EF, Jr., Wright PF, Griffin M R . Rates of hospitalization for respiratory syncytial virus infection among children in medicaid. J Pediatr 2000; 137(6):865-70. 6. Stang P, Brandenburg N , Carter B . The economic burden of respiratory syncytial virus-associated bronchiolitis hospitalizations. Arch Pediatr Adolesc Med 2001; 155(l):95-6. 7. Officer. BCPH. A report on the health of British Columbians: Provincial Health Officer's annual report 1997. Feature report: The health and well-being of British Columbia's children. Victoria, B.C.: Minsitry of Health and Ministry Responsible for Seniors; 1998. 8. Sigurs N , Bjarnason R, Sigurbergsson F, Kjellman B. Respiratory syncytial virus bronchiolitis in infancy is an important risk factor for asthma and allergy at age 7. A m J Respir Crit Care Med 2000; 161(5):1501-7. 9. Sigurs N . Epidemiologic and clinical evidence of a respiratory syncytial virus-reactive airway disease link. A m J Respir Crit Care Med 2001; 163(3 Pt 2):S2-6. 10. Morris JAJ, Blount RE, Savage RE. Recovery of cytopathogenic agent from chimpanzees with coryza. Proc Soc Exp Biol Med 1956; 92:544-550. 11. Chanock R M , Finberg L. Recovery from infants with respiratory illness of a virus related to chimpanzee coryza agent. II. Epidemiological aspects of infection in infants and young children. A m J Hygiene 1957; 66:291-300. 12. Chanock R M , Roizman B , Myers R. Recovery from infants with respiratory illness of a virus related to chimpanzee coryza agent. I. Isolation, properties and characterization. A m J Hygiene 1957; 66:281-290. 13. Collins P, Mcintosh K , Chanock R. Respiratory Syncytial Virus. In: Fields B , Knipe D, Howley P, et al, editors. Fields Virology. Third ed. Philadelphia: Lippincott-Raven; 1996. p. 1313-1351. 128 14. Sullender W M . Respiratory syncytial virus genetic and antigenic diversity. Clin Microbiol Rev 2000; 13(1): 1-15, table of contents. 15. McConnochie K M , Hall C B , Walsh EE , Roghmann K J . Variation in severity of respiratory syncytial virus infections with subtype. J Pediatr 1990; 117(1 Pt l):52-62. 16. Mcintosh ED, De Silva L M , Oates R K . Clinical severity of respiratory syncytial virus group A and B infection in Sydney, Australia. Pediatr Infect Dis J 1993; 12(10):815-9. 17. Wang EE, Law BJ, Stephens D. Pediatric Investigators Collaborative Network on Infections in Canada (PICNIC) prospective study of risk factors and outcomes in patients hospitalized with respiratory syncytial viral lower respiratory tract infection. J Pediatr 1995; 126(2):212-9. 18. Arslanagic E, Matsumoto M , Suzuki K , Nerome K , Tsutsumi H , Hung T. Maturation of respiratory syncytial virus within HEp-2 cell cytoplasm. Acta Virol 1996; 40(4):209-14. 19. Collins PL, Hi l l M G , Camargo E, Grosfeld H , Chanock R M , Murphy BR. Production of infectious human respiratory syncytial virus from cloned cDNA confirms an essential role for the transcription elongation factor from the 5' proximal open reading frame of the M2 mRNA in gene expression and provides a capability for vaccine development. PNAS 1995;92(25):11563-7. 20. Langedijk JP, de Groot B L , Berendsen HJ, van Oirschot JT. Structural homology of the central conserved region of the attachment protein G of respiratory syncytial virus with the fourth subdomain of 55-kDa tumor necrosis factor receptor. Virology 1998; 243(2):293-302. 21. Zhao X , Singh M , Malashkevich V N , Kim PS. Structural characterization of the human respiratory syncytial virus fusion protein core. PNAS 2000; 97(26):14172-7. 22. Garcia J, Garcia-Barreno B, Vivo A , Melero JA. Cytoplasmic inclusions of respiratory syncytial virus-infected cells: formation of inclusion bodies in transfected cells that coexpress the nucleoprotein, the phosphoprotein, and the 22K protein. Virology 1993; 195(l):243-7. 23. Ruuskanen O, Ogra PL. Respiratory syncytial virus. Curr Probl Pediatr 1993; 23(2):50-79. 24. Simoes E A . Respiratory syncytial virus infection. Lancet 1999; 354(9181):847-52. 25. Domachowske JB, Rosenberg HF. Respiratory Syncytial Virus infection: immune response, immunopathogenesis, and treatment. Clin Microbiol Rev 1999; 12:298-309. 26. Hall CB. Respiratory syncytial virus and parainfluenza virus. N Engl J Med 2001; 344(25):1917-28. 129 27. McConnochie K M , Hall CB, Barker WH. Lower respiratory tract illness in the first two years of life: epidemiologic patterns and costs in a suburban pediatric practice. A m J Public Health 1988; 78(l):34-9. 28. Bozzola M , Comoli P, Marchi A , Capra E, Percivalle E, et al. Chromosomal breaks in lymphocytes from respiratory syncytial virus-infected infants. Intl J Med Biol Environ 1999;27(2):147-150. 29. Hull J, Ackerman H , Isles K , Usen S, Pinder M , et al. Unusual haplotypic structure of IL8, a susceptibility locus for a common respiratory virus. A m J Hum Genet 2001; 69(2):413-9. 30. Holberg CJ, Wright A L , Martinez FD, Ray CG, Taussig L M , Lebowitz M D . Risk factors for respiratory syncytial virus-associated lower respiratory illnesses in the first year of life. A m J Epidemiol 1991; 133(11):1135-51. 31. Colocho Zelaya EA, Pettersson C A , Forsgren M , Orvell C, Strannegard O. Respiratory syncytial virus infection in hospitalized patients and healthy children in El Salvador. A m J Trop Med Hyg 1994; 51(5):577-584. 32. La Via W V , Grant SW, Stutman HR, Marks MI. Clinical profile of pediatric patients hospitalized with respiratory syncytial virus infection. Clin Pediatr 1993; 32(8):450-454. 33. Huq F, Rahman M , Nahar N , Alam A , Haque M , et al. Acute lower respiratory tract infection due to virus among hospitalized children in Dhaka, Bangladesh. Rev Infect Dis 1990; 12(Suppl 8):S982-987. 34. Glezen WP, Paredes A , Allison JE, Taber L H , Frank A L . Risk of respiratory syncytial virus infection for infants from low-income families in relationship to age, sex, ethnic group, and maternal antibody level. J Pediatr 1981; 98(5):708-15. 35. Hogg JC. Chronic bronchitis: the role of viruses. Semin Respir Infect 2000; 15(l):32-40. 36. Wohl M E B , Chernick V . State of the art: bronchiolitis. A m Rev Respir Dis 1978; 118(4):759-81. 37. van Schaik S M , Welliver RC, Kimpen JL. Novel pathways in the pathogenesis of respiratory syncytial virus disease. Pediatr Pulmonol 2000; 30(2):131-8. 38. Murphy BR, Prince G A , Walsh EE, K im HW, Parrott RH, et al. Dissociation between serum neutralizing and glycoprotein antibody responses of infants and children who received inactivated respiratory syncytial virus vaccine. J Clin Microbiol 1986; 24(2): 197-202. 39. Kimpen JL. Respiratory syncytial virus immunology. Pediatr Allergy Immunol 1996; 7(9 Suppl):86-90. 130 40. Hogg JC, Williams J, Richardson JB, Macklem PT, Thurlbeck W M . Age as a factor in the distribution of lower-airway conductance and in the pathologic anatomy of obstructive lung disease. N Engl J Med 1970; 282(23): 1283-7. 41. Martinez FD, Morgan WJ, Wright A L , Holberg CJ, Taussig L M . Diminished lung function as a predisposing factor for wheezing respiratory illness in infants. N Engl J Med 1988; 319(17):1112-7. 42. Dargaville PA, South M , McDougall PN. Surfactant abnormalities in infants with severe viral bronchiolitis. Arch Dis Child 1996; 75(2): 133-6. 43. Welliver RC, Sun M , Rinaldo D, Ogra PL. Predictive value of respiratory syncytial virus-specific IgE responses for recurrent wheezing following bronchiolitis. J Pediatr 1986; 109(5):776-80. 44. Welliver RC, Wong DT, Sun M , Middleton E, Jr., Vaughan RS, Ogra PL. The development of respiratory syncytial virus-specific IgE and the release of histamine in nasopharyngeal secretions after infection. N Engl J Med 1981; 305(15):841-6. 45. Welliver RC, Kaul TN, Ogra PL. The appearance of cell-bound IgE in respiratory-tract epithelium after respiratory-syncytial-virus infection. N Engl J Med 1980; 303(21): 1198-202. 46. Welliver RC, Kaul TN, Sun M , Ogra PL. Defective regulation of immune responses in respiratory syncytial virus infection. J Immunol 1984; 133(4): 1925-30. 47. Garofalo R, Kimpen JL, Welliver RC, Ogra PL. Eosinophil degranulation in the respiratory tract during naturally acquired respiratory syncytial virus infection. J Pediatr 1992; 120(l):28-32. 48. Harrison A M , Bonville C A , Rosenberg HF, Domachowske JB. Respiratory syncytical virus-induced chemokine expression in the lower airways: eosinophil recruitment and degranulation. A m J Respir Crit Care Med 1999; 159(6): 1918-24. 49. Ogra PL. From chimpanzee coryza to palivizumab: changing times for respiratory syncytial virus. Pediatr Infect Dis J 2000; 19(8):774-9; discussion 811-3. 50. Law BJ, Wang EE, MacDonald N , McDonald J, Dobson S, et al. Does ribavirin impact on the hospital course of children with respiratory syncytial virus (RSV) infection? An analysis using the pediatric investigators collaborative network on infections in Canada (PICNIC) RSV database. Pediatrics 1997; 99(3):E7. 51. Welliver RC. Respiratory syncytial virus immunoglobulin and monoclonal antibodies in the prevention and treatment of respiratory syncytial virus infection. Semin Perinatal 1998;22(l):87-95. 52. Groothuis JR, Simoes E A , Levin MJ , Hall C B , Long CE, et al. Prophylactic administration of respiratory syncytial virus immune globulin to high-risk infants and 131 young children. The Respiratory Syncytial Virus Immune Globulin Study Group. N Engl J Med 1993; 329(21): 1524-30. 53. Kim HW, Canchola JG, Brandt CD, Pyles G, Chanock R M , et al. Respiratory syncytial virus disease in infants despite prior administration of antigenic inactivated vaccine. A m J Epidemiol 1969; 89(4):422-34. 54. Chin J, Magoffin R L , Shearer L A , Schieble JH, Lennette E H . Field evaluation of a respiratory syncytial virus vaccine and a trivalent parainfluenza virus vaccine in a pediatric population. A m J Epidemiol 1969; 89(4):449-63. 55. Graham BS, Henderson GS, Tang Y W , Lu X , Neuzil K M , Colley DG. Priming immunization determines T helper cytokine mRNA expression patterns in lungs of mice challenged with respiratory syncytial virus. J Immunol 1993; 151(4):2032-40. 56. Hall CB, Powell K R , MacDonald NE, Gala CL, Menegus M E , et al. Respiratory syncytial viral infection in children with compromised immune function. N Engl J Med 1986;315(2):77-81. 57. Englund JA, Sullivan CJ, Jordan M C , Dehner LP, Vercellotti G M , Balfour H H , Jr. Respiratory syncytial virus infection in immunocompromised adults. Ann Intern Med 1988; 109(3):203-8. 58. Whimbey E, Couch R B , Englund JA, Andreeff M , Goodrich J M , et al. Respiratory syncytial virus pneumonia in hospitalized adult patients with leukemia. Clin Infect Dis 1995;21(2):376-9. 59. Bangham CR, Openshaw PJ, Ball L A , King A M , Wertz GW, Askonas B A . Human and murine cytotoxic T cells specific to respiratory syncytial virus recognize the viral nucleoprotein (N), but not the major glycoprotein (G), expressed by vaccinia virus recombinants. J Immunol 1986; 137(12):3973-7. 60. Isaacs D. Viral subunit vaccines. Lancet 1991; 337(8751): 1223-4. 61. Renzi P M , Turgeon JP, Yang JP, Drblik SP, Marcotte JE, et al. Cellular immunity is activated and a TH-2 response is associated with early wheezing in infants after bronchiolitis. J Pediatr 1997; 130(4):584-93. 62. Enhorning G, Duffy L C , Welliver RC. Pulmonary surfactant maintains patency of conducting airways in the rat. A m J Respir Crit Care Med 1995; 151(2 Pt l):554-6. 63. Enhorning G, Holm B A . Disruption of pulmonary surfactant's ability to maintain openness of a narrow tube. J Appl Physiol 1993; 74(6):2922-7. 64. LeVine A M , Gwozdz J, Stark J, Bruno M , Whitsett J, Korfhagen T. Surfactant protein-A enhances respiratory syncytial virus clearance in vivo. J Clin Invest 1999; 103(7): 1015-21. 132 65. Vos GD, Rijtema M N , Blanco CE. Treatment of respiratory failure due to respiratory syncytial virus pneumonia with natural surfactant. Pediatr Pulmonol 1996; 22(6):412-5. 66. Luchetti M , Casiraghi G, Valsecchi R, Galassini E, Marraro G. Porcine-derived surfactant treatment of severe bronchiolitis. Acta Anaesthesiol Scand 1998; 42(7):805-10. 67. Graham BS, Johnson TR, Peebles RS. Immune-mediated disease pathogenesis in respiratory syncytial virus infection. Immunopharmacology 2000; 48(3):237-47. 68. Mosmann TR, Coffman RL. TH1 and TH2 cells: different patterns of lymphokine secretion lead to different functional properties. Annu Rev Immunol 1989; 7:145-73. 69. Roitt I. Essential Immunology. 7th ed. Oxford, UK: Blackwell Scientific Publications; 1991. 70. Woolcock A J , Peat JK. Evidence for the increase in asthma worldwide. Ciba Found Symp 1997; 206:122-34. 71. Steerenberg PA, Van Amsterdam JG, Vandebriel RJ, Vos JG, Van Bree L, Van Loveren H . Environmental and lifestyle factors may act in concert to increase the prevalence of respiratory allergy including asthma. Clin Exp Allergy 1999; 29(10): 1303-8. 72. Strachan DP. Hay fever, hygiene, and household size. B M J 1989; 299(6710):1259-60. 73. Byrd L G , Prince G A . Animal models of respiratory syncytial virus infection. Clin Infect Dis 1997; 25:1363-1368. 74. Piedimonte G, Rodriguez M M , King K A , McLean S, Jiang X . Respiratory syncytial virus upregulates expression of the substance P receptor in rat lungs. A m J Physiol 1999; 277(4 Ptl):L831-40. 75. Wagner M H , Evermann JF, Gaskin J, McNicol K , Small P, Stecenko A A . Subacute effects of respiratory syncytial virus infection on lung function in lambs. Pediatr Pulmonol 1991; ll(l):56-64. 76. Openshaw P. Immunity and immunopathology to respiratory syncytial virus. The mouse model. A m J Respir Crit Care Med 1995; 152:S59-S62. 77. Graham BS, Bunton L A , Wright PF, Karzon DT. Role of T lymphocyte subsets in the pathogenesis of primary infection and rechallenge with respiratory syncytial virus in mice. J Clin Invest 1991; 88(3): 1026-33. 78. Cannon M J , Openshaw PJ, Askonas B A . Cytotoxic T cells clear virus but augment lung pathology in mice infected with respiratory syncytial virus. J Exp Med 1988; 168(3):1163-8. 79. Alwan W H , Record F M , Openshaw PJ. CD4+ T cells clear virus but augment disease in mice infected with respiratory syncytial virus. Comparison with the effects of CD8+ T cells. Clin Exp Immunol 1992; 88(3):527-36. 133 80. Hegele RG, Robinson PJ, Gonzalez S, Hogg JC. Production of acute bronchiolitis in guinea pigs by human respiratory syncytial virus. Eur Respir J 1993; 6(9): 1324-31. 81. Hegele RG, Hayashi S, Bramley A M , Hogg JC. Persistence of respiratory syncytial virus genome and protein after acute bronchiolitis in guinea pigs. Chest 1994; 105(6): 1848-54. 82. Riedel F, Oberdieck B , Streckert HJ, Philippou S, Krusat T, Marek W. Persistence of airway hyperresponsiveness and viral antigen following respiratory syncytial virus bronchiolitis in young guinea-pigs. Eur Respir J 1997; 10(3):639-45. 83. Dakhama A , Kaan P M , Hegele RG. Permissiveness of guinea pig alveolar macrophage subpopulations to acute respiratory syncytial virus infection in vitro. Chest 1998; 114(6):1681-1688. 84. Welliver R. Chemokines, cytokines and inflammatory cells in Respiratory Syncytial Virus infection: similarities to allergic responses. Ped Asthma Allergy Immunol 2000; 14(2):93-100. 85. Gern JE, Busse WW. The role of viral infections in the natural history of asthma. J Allergy Clin Immunol 2000; 106(2):201-12. 86. Zhu Z, Tang W, Ray A , Wu Y , Einarsson O, et al. Rhinovirus stimulation of interleukin-6 in vivo and in vitro. Evidence for nuclear factor kappa B-dependent transcriptional activation. J Clin Invest 1996; 97(2):421-30. 87. Zhu Z, Tang W, Gwaltney J M , Jr., Wu Y , Elias JA. Rhinovirus stimulation of interleukin-8 in vivo and in vitro: role of NF-kappaB. A m J Physiol 1997; 273(4 Pt l):L814-24. 88. Jamaluddin M , Casola A , Garofalo RP, Han Y , Elliott T, et al. The major component of IkappaBalpha proteolysis occurs independently of the proteasome pathway in respiratory syncytial virus-infected pulmonary epithelial cells. J Virol 1998; 72(6):4849-57. 89. Mastronarde JG, Monick M M , Mukaida N , Matsushima K , Hunninghake GW. Activator protein-1 is the preferred transcription factor for cooperative interaction with nuclear factor-kappaB in respiratory syncytial virus-induced interleukin-8 gene expression in airway epithelium. J Infect Dis 1998; 177(5): 1275-81. 90. Chini B A , Fiedler M A , Milligan L, Hopkins T, Stark JM. Essential roles of NF-kappaB and C/EBP in the regulation of intercellular adhesion molecule-1 after respiratory syncytial virus infection of human respiratory epithelial cell cultures. J Virol 1998; 72(2): 1623-6. 91. Becker S, Koren HS, Henke DC. Interleukin-8 expression in normal nasal epithelium and its modulation by infection with respiratory syncytial virus and cytokines tumor necrosis factor, interleukin-1, and interleukin-6. A m J Respir Cell Mol Biol 1993; 8(l):20-7. 134 92. Jiang Z, Kunimoto M , Patel JA. Autocrine regulation and experimental modulation of interleukin-6 expression by human pulmonary epithelial cells infected with respiratory syncytial virus. J Virol 1998; 72(3):2496-9. 93. Kunkel SL, Standiford T, Kasahara K, Strieter R M . Interleukin-8 (IL-8): the major neutrophil chemotactic factor in the lung. Exp Lung Res 1991; 17(1): 17-23. 94. Everard M L , Swarbrick A , Wrightham M , Mclntyre J, Dunkley C, et al. Analysis of cells obtained by bronchial lavage of infants with respiratory syncytial virus infection. Arch Dis Child 1994; 71 (5):428-32. 95. Noah TL, Henderson FW, Wortman IA, Devlin R B , Handy J, et al. Nasal cytokine production in viral acute upper respiratory infection of childhood. J Infect Dis 1995; 171(3):584-92. 96. Arnold R, Werner F, Humbert B , Werchau H , Konig W. Effect of respiratory syncytial virus-antibody complexes on cytokine (IL-8, IL-6, TNF-alpha) release and respiratory burst in human granulocytes. Immunol 1994; 82(2): 184-91. 97. Cromwell O, Hamid Q, Corrigan CJ, Barkans J, Meng Q, et al. Expression and generation of interleukin-8, IL-6 and granulocyte-macrophage colony-stimulating factor by bronchial epithelial cells and enhancement by IL-1 beta and tumour necrosis factor-alpha. Immunol 1992; 77(3):330-7. 98. Noah TL, Becker S. Respiratory syncytial virus-induced cytokine production by a human bronchial epithelial cell line. A m J Physiol 1993; 265(5 Pt l):L472-8. 99. Lotz M . Interleukin-6: a comprehensive review. Cancer Treat Res 1995; 80:209-33. 100. Wilmott RW, Khurana-Hershey G, Stark JM. Current concepts on pulmonary host defense mechanisms in children. Curr Opin Pediatr 2000; 12(3): 187-93. 101. Becker S, Reed W, Henderson FW, Noah TL. RSV infection of human airway epithelial cells causes production of the beta-chemokine RANTES. A m J Physiol 1997; 272(3 Pt l):L512-20. 102. Olszewska-Pazdrak B , Casola A , Saito T, Alam R, Crowe SE, et al. Cell-specific expression of RANTES, MCP-1 , and MIP-1 alpha by lower airway epithelial cells and eosinophils infected with respiratory syncytial virus. J Virol 1998; 72(6):4756-64. 103. Garofalo R, Mei F, Espejo R, Ye G, Haeberle H , et al. Respiratory syncytial virus infection of human respiratory epithelial cells up-regulates class I M H C expression through the induction of IFN-beta and IL-1 alpha. J Immunol 1996; 157(6):2506-13. 104. Wang SZ, Hallsworth PG, Dowling K D , Alpers JH, Bowden JJ, Forsyth K D . Adhesion molecule expression on epithelial cells infected with respiratory syncytial virus. Eur Respir J 2000; 15(2):358-66. 135 105. Kimata H , Yoshida A , Ishioka C, Fujimoto M , Lindley I, Furusho K . RANTES and macrophage inflammatory protein 1 alpha selectively enhance immunoglobulin (IgE) and IgG4 production by human B cells. J Exp Med 1996; 183(5):2397-402. 106. Taub DD, Sayers TJ, Carter CR, Ortaldo JR. Alpha and beta chemokines induce N K cell migration and enhance NK-mediated cytolysis. J Immunol 1995; 155(8):3877-88. 107. Panuska JR, Hertz MI, Taraf H , Villani A , Cirino N M . Respiratory syncytial virus infection of alveolar macrophages in adult transplant patients. A m Rev Resp Dis 1992; 145(4 Pt l):934-9. 108. Midulla F, Huang Y T , Gilbert IA, Cirino N M , McFadden ER, Jr., Panuska JR. Respiratory syncytial virus infection of human cord and adult blood monocytes and alveolar macrophages. A m Rev Resp Dis 1989; 140(3):771-7. 109. Kimpen JL, Garofalo R, Welliver RC, Fujihara K , Ogra PL. A n ultrastructural study of the interaction of human eosinophils with respiratory syncytial virus. Pediatr Allergy Immunol 1996; 7(l):48-53. 110. Jaovisidha P, Peeples M E , Brees A A , Carpenter LR, Moy JN. Respiratory syncytial virus stimulates neutrophil degranulation and chemokine release. J Immunol 1999; 163:2816-2820. 111. Panuska JR, Cirino N M , Midulla F, Despot JE, McFadden ER, Jr., Huang Y T . Productive infection of isolated human alveolar macrophages by respiratory syncytial virus. J Clin Invest 1990; 86(1): 113-9. 112. Becker S, Soukup J, Yankaskas JR. Respiratory syncytial virus infection of human primary nasal and bronchial epithelial cell cultures and bronchoalveolar macrophages. A m J Respir Cell Mol Biol 1992; 6(4):369-74. 113. Cirino N M , Panuska JR, Villani A , Taraf H , Rebert N A , et al. Restricted replication of respiratory syncytial virus in human alveolar macrophages. J Gen Virol 1993; 74(Pt 8):1527-37. 114. Franke-Ullmann G, Pfortner C, Walter P, Steinmuller C, Lohmann-Matthes M L , et al. Alteration of pulmonary macrophage function by respiratory syncytial virus infection in vitro. J Immunol 1995; 154(l):268-80. 115. Trigo E, Liggitt HD, Evermann JF, Breeze RG, Huston L Y , Silflow R. Effect of in vitro inoculation of bovine respiratory syncytial virus on bovine pulmonary alveolar macrophage function. A m J Vet Res 1985; 46(5):1098-103. 116. Becker S, Quay J, Soukup J. Cytokine (tumor necrosis factor, IL-6, and IL-8) production by respiratory syncytial virus-infected human alveolar macrophages. J Immunol 1991; 147:4307-4312. 136 117. Arnold R, Konig B, Galatti H , Werchau H , Konig W. Cytokine (IL-8, IL-6, TNF-alpha) and soluble TNF receptor-I release from human peripheral blood mononuclear cells after respiratory syncytial virus infection. Immunol 1995; 85(3):364-72. 118. Gordon S, Hughes D. Macrophages and their origins. In: Lipscomb M , Russell SW, editors. Lung Macrophages and Dendritic Cells in Health and Disease. New York, N Y : Marcel Dekker Inc.; 1997. p. 3-31. 119. Thomas ED, Ramberg RE, Sale GE, Sparkes RS, Golde DW. Direct evidence for a bone marrow origin of the alveolar macrophage in man. Science 1976; 192(4243): 1016-8. 120. Metcalf D. Transformation of granulocytes to macrophages in bone marrow colonies in vitro. J Cell Physiol 1971; 77(2):277-80. 121. Golde DW, Groopman JE. Production, distribution and fate of monocytes and macrophages. In: Williams WJ, Beutler E, Erslev A J , Lichtman M A , editors. Haematology. 4th Ed. New York: McGraw-Hill; 1990. 122. Cannistra SA, Rambaldi A , Spriggs DR, Herrmann F, Kufe D , Griffin JD. Human granulocyte-macrophage colony-stimulating factor induces expression of the tumor necrosis factor gene by the U937 cell line and by normal human monocytes. J Clin Invest 1987; 79(6): 1720-8. 123. van Furth R. Origin and turnover of monocytes and macrophages. Curr Top Pathol 1989; 79:125-50. 124. Matsushima K , Larsen CG, DuBois GC, Oppenheim J J. Purification and characterization of a novel monocyte chemotactic and activating factor produced by a human myelomonocytic cell line. J Exp Med 1989; 169(4):1485-90. 125. Blusse van Oud Ablas A , van Furth R. Origin, Kinetics, and characteristics of pulmonary macrophages in the normal steady state. J Exp Med 1979; 149(6):1504-18. 126. Bitterman PB, Saltzman L E , Adelberg S, Ferrans V J , Crystal RG. Alveolar macrophage replication. One mechanism for the expansion of the mononuclear phagocyte population in the chronically inflamed lung. J Clin Invest 1984; 74(2):460-9. 127. Nakata K , Akagawa KS , Fukayama M , Hayashi Y , Kadokura M , Tokunaga T. Granulocyte-macrophage colony-stimulating factor promotes the proliferation of human alveolar macrophages in vitro. J Immunol 1991; 147(4): 1266-72. 128. Lohmann-Matthes M L , Steinmuller C, Franke-Ullmann G. Pulmonary macrophages. Eur Respir J 1994; 7(9): 1678-89. 129. Dehring DJ, Wismar B L . Intravascular macrophages in pulmonary capillaries of humans. A m Rev Respir Dis 1989; 139(4): 1027-9. 137 130. Warner A E , Molina R M , Brain JD. Uptake of bloodborne bacteria by pulmonary intravascular macrophages and consequent inflammatory responses in sheep. A m Rev Respir Dis 1987; 136(3):683-90. 131. Sibille Y , Reynolds H Y . Macrophages and polymorphonuclear neutrophils in lung defense and injury. A m Rev Respir Dis 1990; 141(2):471-501. 132. Hance AJ . Pulmonary immune cells in health and disease: dendritic cells and Langerhans' cells. Eur Respir J 1993; 6(8):1213-20. 133. Holt PG, Schon-Hegrad M A , Oliver J. M H C class II antigen-bearing dendritic cells in pulmonary tissues of the rat. Regulation of antigen presentation activity by endogenous macrophage populations. J Exp Med 1988; 167(2):262-74. 134. Lehnert B E , Valdez Y E , Holland L M . Pulmonary macrophages: alveolar and interstitial populations. Exp Lung Res 1985; 9(3-4): 177-90. 135. Warren JS, Kunkel R G , Johnson K J , Ward PA. Comparative 02-. responses of lung macrophages and blood phagocytic cells in the rat. Possible relevance to IgA immune complex induced lung injury. Lab Invest 1987; 57(3):311-20. 136. Holian A, Scheule RK. Alveolar macrophage biology. Hospital Practice 1990;. 137. Sebring RJ, Lehnert BE . Morphometric comparisons of rat alveolar macrophages, pulmonary interstitial macrophages, and blood monocytes. Exp Lung Res 1992; 18(4):479-96. 138. Lehnert BE. Pulmonary and thoracic macrophage subpopulations and clearance of particles from the lung. Environ Health Perspect 1992; 97:17-46. 139. Cohen A B , Cline MJ. The human alveolar macrophage: isolation, cultivation in vitro, and studies of morphologic and functional characteristics. J Clin Invest 1971; 50(7): 1390-8. 140. Shellito J, Kaltreider HB. Heterogeneity of immunologic function among subfractions of normal rat alveolar macrophages. II. Activation as a determinant of functional activity. A m Rev Respir Dis 1985; 131(5):678-83. 141. Holian A , Dauber JH, Diamond MS, Daniele RP. Separation of bronchoalveolar cells from the guinea pig on continuous gradients of Percoll: functional properties of fractionated lung macrophages. J Reticuloendothel Soc 1983; 33(2):157-64. 142. Chandler DB, Fuller WC, Jackson R M , Fulmer JD. Studies of membrane receptors and phagocytosis in subpopulations of rat alveolar macrophages. A m Rev Respir Dis 1986; 133:461-467. 143. Dauber JH, Holian A , Rosemiller M E , Daniele RP. Separation of bronchoalveolar cells from the guinea pig on continuous density gradients of Percoll: morphology and cytochemical properties of fractionated lung macrophages. J Reticuloendothel Soc 1983; 33(2): 119-26. 138 144. Chandler DB, Fulmer JD. Prostaglandin synthesis and release by subpopulations of rat alveolar macrophages. J Immunol 1987; 139(3):893-8. 145. Brannen A L , Chandler DB. Alveolar macrophage subpopulations' responsiveness to chemotactic stimuli. A m J Pathol 1988; 132(l):161-6. 146. Douglas SD, Hassan NF. Morphology of monocytes and macrophages. In: Williams WJ, Beutler E, Erslev A J , Lichtman M A , editors. Haematology. 4th ed. New York: McGraw-Hil l ; 1990. 147. Auger MJ , Ross JA. The biology of the macrophage. In: Lewis CE, McGee JO, editors. The Macrophage. Oxford, U.K. : IRL Press; 1992. 148. Reynolds H Y . Respiratory infections may reflect deficiencies in host defense mechanisms. Dis Mon 1985; 31(2):l-98. 149. Alberts B. Molecular Biology of the Cell. 3rd ed. New York: Garland Publishing Inc.; 1994. 150. Kurt-Jones EA, Popova L , Kwinn L, Haynes L M , Jones LP, et al. Pattern recognition receptors TLR4 and CD 14 mediate response to respiratory syncytial virus. Nat Immunol 2000; 1(5):398-401. 151. Tauber A l , Chernyak L. Metchnikoff and the origins of immunology. Oxford: Oxford University Press; 1991. 152. Griffin F M , Jr., Silverstein SC. Segmental response of the macrophage plasma membrane to a phagocytic stimulus. J Exp Med 1974; 139(2):323-36. 153. Horwitz M A . Formation of a novel phagosome by the Legionnaires' disease bacterium (Legionella pneumophila) in human monocytes. J Exp Med 1983; 158(4): 1319-31. 154. Sung SS, Nelson RS, Silverstein SC. Yeast mannans inhibit binding and phagocytosis of zymosan by mouse peritoneal macrophages. J Cell Biol 1983; 96(l):160-6. 155. Wheater's Functional Histology. 4th ed. New York: Churchill Livingstone; 2000. 156. Nichols B A , Bainton DF, Farquhar M G . Differentiation of monocytes. Origin, nature, and fate of their azurophil granules. J Cell Biol 1971; 50(2):498-515. 157. Johnston RB, Jr., Godzik C A , Cohn Z A . Increased superoxide anion production by immunologically activated and chemically elicited macrophages. J Exp Med 1978; 148(1):115-27. 158. Rouzer C A , Scott W A , Kempe J, Cohn Z A . Prostaglandin synthesis by macrophages requires a specific receptor-ligand interaction. PNAS 1980; 77(7):4279-82. 139 159. Klebanoff SJ. Phagocytic cells: products of oxygen metabolism. In: Gallin JI, Goldstein IM, Snyderman R, editors. Inflammation: Basic principles and clinical correlates. New York: Raven; 1988. 160. Goodman M G , Chenoweth DE, Weigle WO. Induction of interleukin 1 secretion and enhancement of humoral immunity by binding of human C5a to macrophage surface C5a receptors. J Exp Med 1982; 156(3):912-7. 161. Shurin SB, Stossel TP. Complement (C3)-activated phagocytosis by lung macrophages. J Immunol 1978; 120(4): 1305-12. 162. Zeidler R, Eissner G, Meissner P, Uebel S, Tampe R, et al. Downregulation of TAP1 in B lymphocytes by cellular and Epstein-Barr virus-encoded interleukin-10. Blood 1997; 90(6):2390-7. 163. Panuska JR, Merolla R, Rebert N A , Hoffmann SP, Tsivitse P, et al. Respiratory syncytial virus induces interleukin-10 by human alveolar macrophages. Suppression of early cytokine production and implications for incomplete immunity. J Clin Invest 1995; 96(5):2445-53. 164. Kimpen JL. Respiratory syncytial virus and asthma. The role of monocytes. A m J Respir Crit Care Med 2001; 163(3 Pt 2):S7-9. 165. Fels A O , Pawlowski N A , Cramer EB, King TK, Cohn Z A , Scott WA. Human alveolar macrophages produce leukotriene B4. PNAS 1982; 79(24):7866-70. 166. Ming WJ, Bersani L, Mantovani A . Tumor necrosis factor is chemotactic for monocytes and polymorphonuclear leukocytes. J Immunol 1987; 138(5): 1469-74. 167. Schroder J M , Mrowietz U , Morita E, Christophers E. Purification and partial biochemical characterization of a human monocyte-derived, neutrophil-activating peptide that lacks interleukin 1 activity. J Immunol 1987; 139(10):3474-83. 168. Yoshimura T, Matsushima K , Oppenheim JJ, Leonard EJ. Neutrophil chemotactic factor produced by lipopolysaccharide (LPS)-stimulated human blood mononuclear leukocytes: partial characterization and separation from interleukin 1 (IL 1). J Immunol 1987; 139(3):788-93. 169. Wolpe SD, Davatelis G, Sherry B , Beutler B, Hesse DG, et al. Macrophages secrete a novel heparin-binding protein with inflammatory and neutrophil chemokinetic properties. J Exp Med 1988; 167(2):570-81. 170. Goetzl EJ, Pickett WC. Novel structural determinants of the human neutrophil chemotactic activity of leukotriene B. J Exp Med 1981; 153(2):482-7. 171. Henderson WR, Jr. Lipid-derived and other chemical mediators of inflammation in the lung. J Allergy Clin Immunol 1987; 79(4):543-53. 140 172. Dahlen SE, Bjork J, Hedqvist P, Arfors K E , Hammarstrom S, et al. Leukotrienes promote plasma leakage and leukocyte adhesion in postcapillary venules: in vivo effects with relevance to the acute inflammatory response. PNAS 1981; 78(6):3887-91. 173. O'Flaherty JT. Neutrophil degranulation: evidence pertaining to its mediation by the combined effects of leukotriene B4, platelet-activating factor, and 5-HETE. J Cell Physiol 1985; 122(2):229-39. 174. Martin TR, Raugi G, Merritt TL, Henderson WR, Jr. Relative contribution of leukotriene B4 to the neutrophil chemotactic activity produced by the resident human alveolar macrophage. J Clin Invest 1987; 80(4): 1114-24. 175. Carswell E A , Old LJ , Kassel R L , Green S, Fiore N , Williamson B. An endotoxin-induced serum factor that causes necrosis of tumors. PNAS 1975; 72(9):3666-70. 176. Cerami A , Beutler B. The role of cachectin/TNF in endotoxic shock and cachexia. Immunol Today 1988; 9(1):28-31. 177. Nathan CF. Neutrophil activation on biological surfaces. Massive secretion of hydrogen peroxide in response to products of macrophages and lymphocytes. J Clin Invest 1987; 80(6): 1550-60. 178. Shalaby MR, Aggarwal B B , Rinderknecht E, Svedersky LP, Finkle BS, Palladino M A , Jr. Activation of human polymorphonuclear neutrophil functions by interferon-gamma and tumor necrosis factors. J Immunol 1985; 135(3):2069-73. 179. Smith K A , Lachman L B , Oppenheim JJ, Favata MF. The functional relationship of the interleukins. J Exp Med 1980; 151(6):1551-6. 180. Matsuda T, Suematsu S, Kawano M , Yoshizaki K , Tang B, et al. IL-6/BSF2 in normal and abnormal regulation of immune responses. PNAS 1989; 557:466-76; discussion 476-7. 181. Baumann H , Richards C, Gauldie J. Interaction among hepatocyte-stimulating factors, interleukin 1, and glucocorticoids for regulation of acute phase plasma proteins in human hepatoma (HepG2) cells. J Immunol 1987; 139(12):4122-8. 182. Tosato G, Seamon K B , Goldman ND, Sehgal PB, May LT, et al. Monocyte-derived human B-cell growth factor identified as interferon-beta 2 (BSF-2, IL-6). Science 1988; 239(4839):502-4. 183. Mielke V , Bauman JG, Sticherling M , lbs T, Zomershoe A G , et al. Detection of neutrophil-activating peptide NAP/IL-8 and NAP/IL-8 mRNA in human recombinant IL-1 alpha- and human recombinant tumor necrosis factor-alpha-stimulated human dermal fibroblasts. An immunocytochemical and fluorescent in situ hybridization study. J Immunol 1990; 144(1):153-61. 141 184. Larsen CG, Anderson A O , Appella E, Oppenheim JJ, Matsushima K. The neutrophil-activating protein (NAP-1) is also chemotactic for T lymphocytes. Science 1989; 243(4897): 1464-6. 185. Baggiolini M , Dewald B , Moser B. Interleukin-8 and related chemotactic cytokines--C X C and C C chemokines. Adv Immunol 1994; 55:97-179. 186. Driscoll K E . Macrophage inflammatory proteins: biology and role in pulmonary inflammation. Exp Lung Res 1994; 20(6):473-90. 187. Cook DN. The role of MIP-1 alpha in inflammation and hematopoiesis. J Leukoc Biol 1996; 59(l):61-6. 188. Driscoll K E , Hassenbein D G , Carter J, Poynter J, Asquith TN, et al. Macrophage inflammatory proteins 1 and 2: expression by rat alveolar macrophages, fibroblasts, and epithelial cells and in rat lung after mineral dust exposure. A m J Respir Cell Mol Biol 1993; 8(3):311-8. 189. Driscoll K E , Lindenschmidt RC, Maurer JK, Higgins J M , Ridder G. Pulmonary response to silica or titanium dioxide: inflammatory cells, alveolar macrophage-derived cytokines, and histopathology. A m J Respir Cell Mol Biol 1990; 2(4):381-90. 190. Koren HS, Devlin R B , Graham DE, Mann R, McGee MP, et al. Ozone-induced inflammation in the lower airways of human subjects. A m Rev Respir Dis 1989; 139(2):407-15. 191. Alam R, York J, Boyars M , Stafford S, Grant JA, et al. Increased MCP-1, RANTES, and MIP-1 alpha in bronchoalveolar lavage fluid of allergic asthmatic patients. A m J Respir CritCare Med 1996; 153(4 Pt 1): 1398-404. 192. Cruikshank WW, Long A , Tarpy RE, Kornfeld H , Carroll MP, et al. Early identification of interleukin-16 (lymphocyte chemoattractant factor) and macrophage inflammatory protein 1 alpha (MIP1 alpha) in bronchoalveolar lavage fluid of antigen-challenged asthmatics. A m J Respir Cell Mo l Biol 1995; 13(6):738-47. 193. Becker S, Soukup JM. Exposure to urban air particulates alters the macrophage-mediated inflammatory response to respiratory viral infection. J Toxicol Environ Health 1999; 57(7):445-57. 194. Garofalo RP, Parti J, Hintz K A , Hi l l V , Ogra PL, Welliver RC. Macrophage inflammatory protein-1 alpha (not T helper type 2 cytokines) is associated with severe forms of respiratory syncytial virus bronchiolitis. J Infect Dis 2001; 184(4):393-9. 195. Sibille Y , Merrill WW, Naegel GP, Care SB, Cooper J A , Jr., Reynolds H Y . Human alveolar macrophages release a factor that inhibits phagocyte function. A m J Respir Cell Mol Biol 1989; 1(5):407-16. 142 196. Thomassen MJ , Buhrow LT, Connors M J , Kaneko FT, Erzurum SC, Kavuru MS. Nitric oxide inhibits inflammatory cytokine production by human alveolar macrophages. A m J Respir Cell Mol Biol 1997; 17(3):279-83. 197. Canadian Council on Animal Care (CCAC). Ottawa: C C A C ; 1980. 198. Bendayan M , Zollinger M . Ultrastructural localization of antigenic sites on osmium fixed tissues applying the protein A-gold technique. J Histochem Cytochem 1983; 31(l):101-9. 199. Nakstad B , Lyberg T, Skjorten F, Boye NP. Subpopulations of human lung alveolar macrophages: ultrastructural features. Ultrastructural Pathology 1989; 13(1):1-13. 200. Phillips MJ , Blendis L M , Poucell S, offterson J, Petric M , et al. Syncytial giant-cell hepatitis. Sporadic hepatitis with distinctive pathological features, a severe clinical course, and paramyxoviral features. N Engl J Med 1991; 324(7):455-60. 201. Ferro TJ, Kern JA, Elias JA, Kamoun M , Daniele RP, Rossman M D . Alveolar macrophages, blood monocytes, and density-fractionated alveolar macrophages differ in their ability to promote lymphocyte proliferation to mitogen and antigen. A m Rev Respir Dis 1987; 135:682-687. 202. Hirsch MS, Zisman B , Allison A C . Macrophages and age-dependent resistance to Herpes simplex virus in mice. J Immunol 1970; 104(5): 1160-5. 203. Mogensen SC. Macrophages and age-dependent resistance to hepatitis induced by herpes simplex virus type 2 in mice. Infect Immun 1978; 19(l):46-50. 204. Turner GS, Ballard R. Interaction of mouse peritoneal macrophages with fixed rabies virus in vivo and in vitro. J Gen Virol 1976; 30(2):223-31. 205. Mogensen SC. Role of macrophages in natural resistance to virus infections. Microbiol Rev 1979; 43(1): 1-26. 206. Prince G A , Porter DD. The pathogenesis of respiratory syncytial virus infection in infant ferrets. A m J Pathol 1976; 82:339-52. 207. Hegele RG, Robinson PJ, Hogg JC. Age susceptibility of guinea pigs to acute bronchiolitis induced by human respiratory syncytial virus. A m Rev Respir Dis 1992; 145:A434. 208. Wagner JE, Manning PJ, editors. The biology of the guinea pig. London, UK: Academic Press Inc.; 1976. 209. Welliver RC. Immunology of respiratory syncytial virus infection: eosinophils, cytokines, chemokines and asthma. Pediatr Infect Dis J 2000; 19(8):780-3; discussion 784-5; 811-3. 210. McMillan E M , King G M , Adamson IYR. Sex hormones influence growth and surfactant production in fetal lung explants. Exp Lung Res 1988; 15:167-179. 143 211. Adamson IYR, King G M . Sex-related differences in cellular composition and surfactant synthesis of developing fetal rat lungs. A m Rev Respir Dis 1984; 129:130-134. 212. Mestan J, Digel W, Mittnacht S, Hillen H , Blohm D, et al. Antiviral effects of recombinant tumour necrosis factor in vitro. Nature 1986; 323(6091):816-9. 213. Wong G H , Goeddel DV. Tumour necrosis factors alpha and beta inhibit virus replication and synergize with interferons. Nature 1986; 323(6091):819-22. 214. Horiuchi T, Tsang S, Schellenberg RR. Release of Tumor Necrosis Factor alpha from sensitized guinea pig lung by allergen. A m J Respir Crit Care Med 1994; 149:A769. 215. von Mutius E, Martinez F, Fritzsch C, Nicolai T, Roell G, Thiemann H . Prevalence of asthma and atopy in two areas of West and East Germany. A m J Respir Crit Care Med 1994; 149:358-364. 216. Weiland SK, von Mutius E, Hirsch T, Duhme H, Fritzsch C, et al. Prevalence of respiratory and atopic disorders among children in the East and West of Germany five years after unification. Eur Respir J 1999; 14(4):862-70. 217. Pope C A , 3rd, Dockery DW. Acute health effects of PMIO pollution on symptomatic and asymptomatic children. A m Rev Respir Dis 1992; 145(5): 1123-8. 218. French JG, Lowrimore G, Nelson WC, Finklea JF, English T, Hertz M . The effect of sulfur dioxide and suspended sulfates on acute respiratory disease. Arch Environ Health 1973;27(3):129-33. 219. Speizer FE, Ferris B , Jr., Bishop Y M , Spengler J. Respiratory disease rates and pulmonary function in children associated with N 0 2 exposure. A m Rev Respir Dis 1980; 121(1):3-10. 220. Ball T M , Castro-Rodriguez JA, Griffith K A , Holberg CJ, Martinez FD, Wright A L . Siblings, day-care attendance, and the risk of asthma and wheezing during childhood. N Engl J Med 2000; 343(8):538-43. 221. Butland B K , Strachan DP, Lewis S, Bynner J, Butler N , Britton J. Investigation into the increase in hay fever and eczema at age 16 observed between the 1958 and 1970 British birth cohorts. B M J 1997; 315(7110):717-21. 222. Strachan DP, Anderson HR, Limb ES, O'Neill A , Wells N . A national survey of asthma prevalence, severity, and treatment in Great Britain. Arch Dis Child 1994; 70(3): 174-8. 223. Shaheen SO, Aaby P, Hall AJ , Barker DJ, Heyes CB, et al. Measles and atopy in Guinea-Bissau. Lancet 1996; 347(9018): 1792-6. 224. Bodner C, Godden D, Seaton A . Family size, childhood infections and atopic diseases. The Aberdeen W H E A S E Group. Thorax 1998; 53(l):28-32. 144 225. Shirakawa T, Enomoto T, Shimazu S, Hopkin JM. The inverse association between tuberculin responses and atopic disorder. Science 1997; 275(5296):77-9. 226. Matricardi P M , Rosmini F, Ferrigno L, Nisini R, Rapicetta M , et al. Cross sectional retrospective study of prevalence of atopy among Italian military students with antibodies against hepatitis A virus. B M J 1997; 314(7086):999-1003. 227. Paunio M , Heinonen OP, Virtanen M , Leinikki P, Patja A , Peltola H. Measles history and atopic diseases: a population-based cross-sectional study. J A M A 2000; 283(3):343-6. 228. Gern JE, Weiss ST. Protection against atopic diseases by measles\u00E2\u0080\u0094a rash conclusion? J A M A 2000; 283(3):394-5. 229. Strannegard IL, Larsson LO, Wennergren G, Strannegard O. Prevalence of allergy in children in relation to prior B C G vaccination and infection with atypical mycobacteria. Allergy 1998; 53(3):249-54. 230. Dockery DW, Pope C A , 3rd. Acute respiratory effects of particulate air pollution. Annu Rev Public Health 1994; 15:107-32. 231. Schwartz J. Air pollution and daily mortality: a review and meta analysis. Environ Res 1994; 64(l):36-52. 232. von Mutius E, Fritzsch C, Weiland SK, Roll G, Magnussen H . Prevalence of asthma and allergic disorders among children in united Germany: a descriptive comparison. B M J 1992; 305(6866): 1395-9. 233. Nowak D, Heinrich J, Jorres R, Wassmer G, Berger J, et al. Prevalence of respiratory symptoms, bronchial hyperresponsiveness and atopy among adults: West and East Germany. Eur Respir J 1996; 9(12):2541-52. 234. Braback L , Breborowicz A , Julge K , Knutsson A , Riikjarv M A , et al. Risk factors for respiratory symptoms and atopic sensitisation in the Baltic area. Arch Dis Child 1995; 72(6):487-93. 235. Braback L, Breborowicz A , Dreborg S, Knutsson A , Pieklik H, Bjorksten B. Atopic sensitization and respiratory symptoms among Polish and Swedish school children. Clin Exp Allergy 1994; 24(9):826-35. 236. Becker S, Soukup J, Gilmour M , Devlin R. Stimulation of human and rat alveolar macrophages by urban air particulates: effects on oxidant radical generation and cytokine production. Toxicol Appl Pharmacol 1996; 141:637-648. 237. Kaan P M , Hegele RG. Effect of PM10 on cytokine production by guinea pig alveolar macrophages. A m J Respir Crit Care Med 2001; 163(5):A362. 238. Kaan P M , Hegele RG. Effect of PM10 on cell viability and cytokine production by subpopulations of guinea pig alveolar macrophages. A m J Respir Crit Care Med 2001; 163(6):A188. 145 239. Vincent R, Bjarnason SG, Adamson IY, Hedgecock C, Kumarathasan P, et al. Acute pulmonary toxicity of urban particulate matter and ozone. A m J Pathol 1997; 151(6): 1563-70. 240. Goldsmith C-A, Imrich A , Danaee H , Kobzik L , Ning Y Y . Analysis of air pollution particulate-mediated oxidant stress in alveolar macrophages. J Toxicol Environ Health 1998;54:529-545. 241. Ghio AJ , Carter JD, Richards JH, Brighton LE , Lay JC, Devlin R B . Disruption of normal iron homeostasis after bronchial instillation of an iron-containing particle. A m J Physiol 1998;274:L396-L403. 242. Kennedy T, Ghio A J , Reed W, Samet J M , Zagorski J, et al. Copper-dependent inflammation and nuclear factor-kappa B activation by particulate air pollution. A m J Respir Cell Mol Biol 1998; 19:366-378. 243. Carter JD, Ghio AJ , Reed W, Devlin RB. Cytokine production by human airway epithelial cells after exposure to an air pollution particle is metal-dependent. Toxicol Appl Pharmacol 1997; 146:180-188. 244. Hallden G, Skold C M , Eklund A , Forslid J, Hed J. Quenching of intracellular autofluorescence in alveolar macrophages permits analysis of fluorochrome labelled surface antigens by flow cytofluorometry. J Immunol Methods 1991; 142(2):207-14. 245. Lundborg M , Johard U , Lastbom L, Gerde P, Camner P. Human alveolar macrophage phagocytic function is impaired by aggregates of ultrafine carbon particles. Environ Res 2001;86(3):244-53. 246. Lundborg M , Johansson A , Lastbom L, Camner P. Ingested aggregates of ultrafine carbon particles and interferon-gamma impair rat alveolar macrophage function. Environ Res 1999;81(4):309-15. 247. Renwick L C , Donaldson K , Clouter A . Impairment of alveolar macrophage phagocytosis by ultrafine particles. Toxicol Appl Pharmacol 2001; 172(2): 119-27. 248. L i X Y , Gilmour PS, Donaldson K , MacNee W. In vivo and in vitro proinflammatory effects of particulate air pollution (PM10). Environ Health Perspect 1997; 105 Suppl 5:1279-83. 249. Tsutsumi H , Takeuchi R, Ohsaki M , Seki K , Chiba S. Respiratory syncytial virus infection of human respiratory epithelial cells enhances inducible nitric oxide synthase gene expression. J Leukoc Biol 1999; 66(1):99-104. 250. Asano K , Chee C B , Gaston B , Li l ly C M , Gerard C, et al. Constitutive and inducible nitric oxide synthase gene expression, regulation, and activity in human lung epithelial cells. PNAS 1994; 91(21):10089-93. 146 251. Majano PL, Garcia-Monzon C, Lopez-Cabrera M , Lara-Pezzi E, Fernandez-Ruiz E, et al. Inducible nitric oxide synthase expression in chronic viral hepatitis. Evidence for a virus-induced gene upregulation. J Clin Invest 1998; 101 (7): 1343-52. 252. Reiss CS, Komatsu T. Does nitric oxide play a critical role in viral infections? J Virol 1998; 72(6):4547-51. 253. Lopez-Guerrero J A , Carrasco L . Effect of nitric oxide on poliovirus infection of two human cell lines. J Virol 1998; 72(3):2538-40. 254. Harris N , Buller R M , Karupiah G. Gamma interferon-induced, nitric oxide-mediated inhibition of vaccinia virus replication. J Virol 1995; 69(2):910-5. 255. Zaragoza C, Ocampo CJ, Saura M , McMillan A , Lowenstein CJ. Nitric oxide inhibition of coxsackievirus replication in vitro. J Clin Invest 1997; 100(7): 1760-7. 256. Sanders SP, Siekierski ES, Porter JD, Richards S M , Proud D. Nitric oxide inhibits rhinovirus-induced cytokine production and viral replication in a human respiratory epithelial cell line. J Virol 1998; 72(2):934-42. 257. Geist LJ , Hunninghake GW. Role of viruses in cytokine induction from lung cells. In: Nelson S, Martin TR, editors. Cytokines in pulmonary disease. Infection and inflammation. New York: Marcel Dekker, Inc.; 2000. p. 293-305. 258. Hamilton TA, Ohmori Y , Tebo JM, Kishore R. Regulation of macrophage gene expression by pro- and anti-inflammatory cytokines. Pathobiology 1999; 67(5-6):241-4. 259. Mastronarde JG, He B , Monick M M , Mukaida N , Matsushima K , Hunninghake GW. Induction of interleukin (IL)-8 gene expression by respiratory syncytial virus involves activation of nuclear factor (NF)-kappa B and NF-IL-6. J Infect Dis 1996; 174(2):262-7. 260. Bitko V , Velazquez A , Yang L , Yang Y C , Barik S. Transcriptional induction of multiple cytokines by human respiratory syncytial virus requires activation of NF-kappa B and is inhibited by sodium salicylate and aspirin. Virology 1997; 232(2):369-78. 261. Thomas L H , Friedland JS, Sharland M , Becker S. Respiratory syncytial virus-induced RANTES production from human bronchial epithelial cells is dependent on nuclear factor-kappa B nuclear binding and is inhibited by adenovirus-mediated expression of inhibitor of kappa B alpha. J Immunol 1998; 161(2):1007-16. 262. Fiedler M A , Wernke-Dollries K , Stark JM. Inhibition of viral replication reverses respiratory syncytial virus-induced NF-kappaB activation and interleukin-8 gene expression in A549 cells. J Virol 1996; 70(12):9079-82. 263. Sen R, Baltimore D. Inducibility of kappa immunoglobulin enhancer-binding protein Nf-kappa B by a posttranslational mechanism. Cell 1986; 47(6):921-8. 264. Ghosh S, Baltimore D. Activation in vitro of NF-kappa B by phosphorylation of its inhibitor I kappa B. Nature 1990; 344(6267):678-82. 147 265. Baeuerle PA, Henkel T. Function and activation of NF-kappa B in the immune system. Annu Rev Immunol 1994; 12:141-79. 266. Foo SY, Nolan GP. NF-kappaB to the rescue: RELs, apoptosis and cellular transformation. Trends Genet 1999; 15(6):229-35. 267. Baeuerle PA, Baltimore D. NF-kappa B: ten years after. Cell 1996; 87(1): 13-20. 268. Kunsch C, Lang R K , Rosen C A , Shannon MF. Synergistic transcriptional activation of the IL-8 gene by NF-kappa B p65 (RelA) andNF-IL-6. J Immunol 1994; 153(1): 153-64. 269. Collart M A , Baeuerle P, Vassalli P. Regulation of tumor necrosis factor alpha transcription in macrophages: involvement of four kappa B-like motifs and of constitutive and inducible forms of NF-kappa B. Mo l Cell Biol 1990; 10(4): 1498-506. 270. Drouet C, Shakhov A N , Jongeneel C V . Enhancers and transcription factors controlling the inducibility of the tumor necrosis factor-alpha promoter in primary macrophages. J Immunol 1991; 147(5):1694-700. 271. Libermann TA, Baltimore D. Activation of interleukin-6 gene expression through the NF-kappa B transcription factor. Mo l Cell Biol 1990; 10(5):2327-34. 272. Matsusaka T, Fujikawa K , Nishio Y , Mukaida N , Matsushima K , et al. Transcription factors NF-IL6 and NF-kappa B synergistically activate transcription of the inflammatory cytokines, interleukin 6 and interleukin 8. PNAS 1993; 90(21):10193-7. 273. Baichwal V R , Baeuerle PA. Activate NF-kappa B or die? Curr Biol 1997; 7(2):R94-6. 274. Marianneau P, Cardona A , Edelman L, Deubel V , Despres P. Dengue virus replication in human hepatoma cells activates NF-kappaB which in turn induces apoptotic cell death. J Virol 1997; 71(4):3244-9. 275. Garofalo R, Sabry M , Jamaluddin M , Y u RK, Casola A , et al. Transcriptional activation of the interleukin-8 gene by respiratory syncytial virus infection in alveolar epithelial cells: nuclear translocation of the RelA transcription factor as a mechanism producing airway mucosal inflammation. J Virol 1996; 70(12):8773-81. 276. Mastronarde JG, Monick M M , Gross TJ, Hunninghake GW. Amiloride inhibits cytokine production in epithelium infected with respiratory syncytial virus. A m J Physiol 1996; 271(2Ptl):L201-7. 277. Jamaluddin M , Garofalo R, Ogra PL, Brasier AR. Inducible translational regulation of the NF-IL6 transcription factor by respiratory syncytial virus infection in pulmonary epithelial cells. J Virol 1996; 70(3): 1554-63. 278. Mukaida N , Okamoto S, Ishikawa Y , Matsushima K. Molecular mechanism of interleukin-8 gene expression. J Leukoc Biol 1994; 56(5):554-8. 148 279. Casola A , Garofalo RP, Jamaluddin M , Vlahopoulos S, Brasier AR. Requirement of a novel upstream response element in respiratory syncytial virus-induced IL-8 gene expression. J Immunol 2000; 164(11):5944-51. 280. Bitko V , Barik S. Persistent activation of RelA by respiratory syncytial virus involves protein kinase C, underphosphorylated IkappaBbeta, and sequestration of protein phosphatase 2A by the viral phosphoprotein. J Virol 1998; 72(7):5610-8. 281. Baeuerle PA, Baichwal V R . NF-kappa B as a frequent target for immunosuppressive and anti-inflammatory molecules. Adv Immunol 1997; 65:111-37. 282. Kopp EB, Ghosh S. NF-kappa B and rel proteins in innate immunity. Adv Immunol 1995;58:1-27. 283. Delhase M , L i N , Karin M . Kinase regulation in inflammatory response. Nature 2000; 406(6794):367-8. 284. Ozes ON, Akca H , Mayo LD, Gustin JA, Maehama T, et al. A phosphatidylinositol 3-kinase/Akt/mTOR pathway mediates and PTEN antagonizes tumor necrosis factor inhibition of insulin signaling through insulin receptor substrate-1. PNAS 2001; 98(8):4640-5. 285. Karin M , Delhase M . The I kappa B kinase (IKK) and NF-kappa B: key elements of proinflammatory signalling. Semin Immunol 2000; 12(l):85-98. 286. Muzio M , Natoli G, Saccani S, Levrero M , Mantovani A . The human toll signaling pathway: divergence of nuclear factor kappaB and JNK/SAPK activation upstream of tumor necrosis factor receptor-associated factor 6 (TRAF6). J Exp Med 1998; 187(12):2097-101. 287. Benn J, Schneider RJ. Hepatitis B virus HBx protein activates Ras-GTP complex formation and establishes a Ras, Raf, M A P kinase signaling cascade. PNAS 1994; 91(22):10350-4. 288. Doria M , Klein N , Lucito R, Schneider RJ. The hepatitis B virus HBx protein is a dual specificity cytoplasmic activator of Ras and nuclear activator of transcription factors. E M B O J 1995; 14(19):4747-57. 289. Bruder JT, Kovesdi I. Adenovirus infection stimulates the Raf/MAPK signaling pathway and induces interleukin-8 expression. J Virol 1997; 71(l):398-404. 290. Chen W, Monick M M , Carter A B , Hunninghake GW. Activation of ERK2 by respiratory syncytial virus in A549 cells is linked to the production of interleukin 8. Exp Lung Res 2000; 26(l):13-26. 291. Lange-Carter C A , Pleiman C M , Gardner A M , Blumer KJ , Johnson GL. A divergence in the M A P kinase regulatory network defined by M E K kinase and Raf. Science 1993; 260(5106):315-9. 149 292. Moodie SA, Willumsen B M , Weber M J , Wolfman A. Complexes of Ras.GTP with Raf-1 and mitogen-activated protein kinase kinase. Science 1993; 260(5114):1658-61. 293. Denhardt DT. Signal-transducing protein phosphorylation cascades mediated by Ras/Rho proteins in the mammalian cell: the potential for multiplex signaling. Biochem J 1996; 318(Pt3):729-47. 294. Lewis TS, Shapiro PS, Ahn N G . Signal transduction through M A P kinase cascades. Adv Cancer Res 1998; 74:49-139. 295. Monick M , Staber J, Thomas K , Hunninghake G. Respiratory syncytial virus infection results in activation of multiple protein kinase C isoforms leading to activation of mitogen-activated protein kinase. J Immunol 2001; 166(4):2681-7. 296. Duke RC. Apoptosis in cell-mediated immunity. In: Tomei L D , Cope FO, editors. Apoptosis: the molecular basis of cell death. Cold Spring Harbor, N Y : Cold Spring Harbor Laboratory; 1991. p. 209-226. 297. Gougeon M L , Montagnier L. Apoptosis in AIDS. Science 1993; 260(5112): 1269-70. 298. Terai C, Kornbluth RS, Pauza CD, Richman DD, Carson DA. Apoptosis as a mechanism of cell death in cultured T lymphoblasts acutely infected with HIV-1. J Clin Invest 1991; 87(5):1710-5. 299. Hinshaw V S , Olsen CW, Dybdahl-Sissoko N , Evans D. Apoptosis: a mechanism of cell killing by influenza A and B viruses. J Virol 1994; 68(6):3667-73. 300. Takizawa T, Matsukawa S, Higuchi Y , Nakamura S, Nakanishi Y , Fukuda R. Induction of programmed cell death (apoptosis) by influenza virus infection in tissue culture cells. J Gen Virol 1993; 74(Pt ll):2347-55. 301. Teodoro JG, Branton PE. Regulation of apoptosis by viral gene products. J Virol 1997; 71(3):1739-46. 302. McFadden G. Even viruses can learn to cope with stress. Science 1998; 279(5347):40-l. 303. Henderson S, Huen D, Rowe M , Dawson C, Johnson G, Rickinson A . Epstein-Barr virus-coded BHRF1 protein, a viral homologue of Bcl-2, protects human B cells from programmed cell death. PNAS 1993; 90(18):8479-83. 304. Rao L , Debbas M , Sabbatini P, Hockenbery D, Korsmeyer S, White E. The adenovirus E l A proteins induce apoptosis, which is inhibited by the E1B 19-kDa and Bcl-2 proteins. PNAS 1992; 89(16):7742-6. 305. Schreiber M , Sedger L , McFadden G. Distinct domains of M-T2, the myxoma virus tumor necrosis factor (TNF) receptor homolog, mediate extracellular TNF binding and intracellular apoptosis inhibition. J Virol 1997; 71(3):2171-81. 150 306. O'Donnell D R, Milligan L, Stark JM. Induction of CD95 (Fas) and apoptosis in respiratory epithelial cell cultures following respiratory syncytial virus infection. Virology 1999; 257(l):198-207. 307. Bitko V , Barik S. A n endoplasmic reticulum-specific stress-activated caspase (caspase-12) is implicated in the apoptosis of A549 epithelial cells by respiratory syncytial virus. J Cell Biochem 2001; 80(3):441-54. 308. Nakagawa T, Zhu H , Morishima N , L i E, X u J, et al. Caspase-12 mediates endoplasmic-reticulum-specific apoptosis and cytotoxicity by amyloid-beta. Nature 2000; 403(6765):98-103. 309. Takeuchi R, Tsutsumi H , Osaki M , Haseyama K , Mizue N , Chiba S. Respiratory syncytial virus infection of human alveolar epithelial cells enhances interferon regulatory factor 1 and interleukin-1 beta-converting enzyme gene expression but does not cause apoptosis. J Virol 1998; 72(5):4498-502. 310. Cohen G M . Caspases: the executioners of apoptosis. Biochem J 1997; 326(Pt 1):1-16. 311. Domachowske JB, Bonville C A , Mortelliti A J , Colella CB, K i m U , Rosenberg HF. Respiratory syncytial virus infection induces expression of the anti-apoptosis gene IEX-1L in human respiratory epithelial cells. J Infect Dis 2000; 181(3):824-30. 312. Kondratyev A D , Chung K N , Jung M O . Identification and characterization of a radiation-inducible glycosylated human early-response gene. Cancer Res 1996; 56(7): 1498-502. 313. Wu M X , Ao Z, Prasad K V , Wu R, Schlossman SF. IEX-1 L , an apoptosis inhibitor involved in NF-kappaB-mediated cell survival. Science 1998; 281(5379):998-1001. 314. Krilov LR, McCloskey TW, Harkness SH, Pontrelli L, Pahwa S. Alterations in apoptosis of cord and adult peripheral blood mononuclear cells induced by in vitro infection with respiratory syncytial virus. J Infect Dis 2000; 181(l):349-53. 315. Laemmli U K . Cleavage of structural proteins during the assembly of the head of bacteriophage T4. Nature 1970; 227(259):680-5. 316. Lowry OH. Protein measurement with the Folin Phenol reagent. J Biol Chem 1951; 193:265-275. 317. Peterson GL. Review of the Folin phenol protein quantitation method of Lowry, Rosebrough, Fair and Randall. Anal Biochem 1979; 100(2):201-20. 318. Riches DW, Chan ED, Winston BW. TNF-alpha-induced regulation and signalling in macrophages. Immunobiol 1996; 195(4-5):477-90. 319. Frodin M , Gammeltoft S. Role and regulation of 90 kDa ribosomal S6 kinase (RSK) in signal transduction. Mol Cell Endocrinol 1999; 151(l-2):65-77. 151 320. Shimamura A , Ballif B A , Richards SA, Blenis J. Rskl mediates a M E K - M A P kinase cell survival signal. Curr Biol 2000; 10(3): 127-35. 321. Schouten GJ, Vertegaal A C , Whiteside ST, Israel A , Toebes M , et al. IkappaB alpha is a target for the mitogen-activated 90 kDa ribosomal S6 kinase. E M B O J 1997; 16(ll):3133-44. 322. Ghoda L, Lin X , Greene WC. The 90-kDa ribosomal S6 kinase (pp90rsk) phosphorylates the N-terminal regulatory domain of IkappaBalpha and stimulates its degradation in vitro. J Biol Chem 1997; 272(34):21281-8. 323. Luo H , Yanagawa B , Zhang J, Luo Z, Zhang M , et al. Coxsackievirus B3 replication is reduced by inhibition of the extracellular signal-regulated kinase (ERK) signaling pathway. J Virol 2002; [In Press]. 324. Datta SR, Brunet A , Greenberg M E . Cellular survival: a play in three Akts. Genes Dev 1999; 13(22):2905-27. 325. Fruman DA, Meyers RE, Cantley L C . Phosphoinositide kinases. Annu Rev Biochem 1998;67:481-507. 326. Cross D A , Alessi DR, Cohen P, Andjelkovich M , Hemmings B A . Inhibition of glycogen synthase kinase-3 by insulin mediated by protein kinase B. Nature 1995; 378(6559):785-9. 327. Kulik G, Klippel A , Weber MJ . Antiapoptotic signalling by the insulin-like growth factor I receptor, phosphatidylinositol 3-kinase, and Akt. Mol Cell Biol 1997; 17(3): 1595-606. 328. Dudek H , Datta SR, Franke TF, Birnbaum M J , Yao R, et al. Regulation of neuronal survival by the serine-threonine protein kinase Akt. Science 1997; 275(5300):661-5. 329. Kennedy SG, Wagner A J , Conzen SD, Jordan J, Bellacosa A , et al. The PI 3-kinase/Akt signaling pathway delivers an anti-apoptotic signal. Genes Dev 1997; 11(6):701-13. 330. Staal SP. Molecular cloning of the akt oncogene and its human homologues AKT1 and AKT2: amplification of AKT1 in a primary human gastric adenocarcinoma. PNAS 1987; 84(14):5034-7. 331. Miwa W, Yasuda J, Murakami Y , Yashima K , Sugano K , et al. Isolation of D N A sequences amplified at chromosome 19ql3.1-ql3.2 including the A K T 2 locus in human pancreatic cancer. Biochem Biophys Res Commun 1996; 225(3):968-74. 332. Cheng JQ, Ruggeri B , Klein W M , Sonoda G, Altomare D A , et al. Amplification of AKT2 in human pancreatic cells and inhibition of A K T 2 expression and tumorigenicity by antisense RNA. PNAS 1996; 93(8):3636-41. 333. Bellacosa A , de Feo D, Godwin A K , Bell DW, Cheng JQ, et al. Molecular alterations of the A K T 2 oncogene in ovarian and breast carcinomas. Int J Cancer 1995; 64(4):280-5. 152 334. Carson JP, Kulik G, Weber MJ. Antiapoptotic signaling in LNCaP prostate cancer cells: a survival signaling pathway independent of phosphatidylinositol 3'-kinase and Akt/protein kinase B. Cancer Res 1999; 59(7): 1449-53. 335. Khwaja A , Rodriguez-Viciana P, Wennstrom S, Warne PH, Downward J. Matrix adhesion and Ras transformation both activate a phosphoinositide 3-OH kinase and protein kinase B/Akt cellular survival pathway. E M B O J 1997; 16(10):2783-93. 336. Bellacosa A , Testa JR, Staal SP, Tsichlis PN. A retroviral oncogene, akt, encoding a serine-threonine kinase containing an SH2-like region. Science 1991; 254(5029):274-7. 337. Franke TF, Yang SI, Chan TO, Datta K , Kazlauskas A , et al. The protein kinase encoded by the Akt proto-oncogene is a target of the PDGF-activated phosphatidylinositol 3-kinase. Cell 1995; 81(5):727-36. 338. Burgering B M , Coffer PJ. Protein kinase B (c-Akt) in phosphatidylinositol-3-OH kinase signal transduction. Nature 1995; 376(6541 ):599-602. 339. Franke TF, Kaplan DR, Cantley L C , Toker A . Direct regulation of the Akt proto-oncogene product by phosphatidylinositol-3,4-bisphosphate. Science 1997; 275(5300):665-8. 340. Alessi DR, James SR, Downes CP, Holmes A B , Gaffney PR, et al. Characterization of a 3-phosphoinositide-dependent protein kinase which phosphorylates and activates protein kinase Balpha. CurrBiol 1997; 7(4):261-9. 341. Stokoe D, Stephens LR, Copeland T, Gaffney PR, Reese C B , et al. Dual role of phosphatidylinositol-3,4,5-trisphosphate in the activation of protein kinase B . Science 1997;277(5325):567-70. 342. Toker A , Newton A C . Cellular signaling: pivoting around PDK-1. Cell 2000; 103(2): 185-8. 343. Deprez J, Vertommen D, Alessi DR, Hue L, Rider M H . Phosphorylation and activation of heart 6-phosphofructo-2-kinase by protein kinase B and other protein kinases of the insulin signaling cascades. J Biol Chem 1997; 272(28): 17269-75. 344. Datta SR, Dudek H , Tao X , Masters S, Fu H , et al. Akt phosphorylation of B A D couples survival signals to the cell-intrinsic death machinery. Cell 1997; 91(2):231-41. 345. del Peso L, Gonzalez-Garcia M , Page C, Herrera R, Nunez G. Interleukin-3-induced phosphorylation of B A D through the protein kinase Akt. Science 1997; 278(5338):687-9. 346. Alessi DR, Caudwell FB, Andjelkovic M , Hemmings B A , Cohen P. Molecular basis for the substrate specificity of protein kinase B ; comparison with M A P K A P kinase-1 and p70 S6 kinase. FEBS Lett 1996; 399(3):333-8. 347. Kandel ES, Hay N . The regulation and activities of the multifunctional serine/threonine kinase Akt /PKB. Exp Cell Res 1999; 253(l):210-29. 153 348. Cardone M H , Roy N , Stennicke HR, Salvesen GS, Franke TF, et al. Regulation of cell death protease caspase-9 by phosphorylation. Science 1998; 282(5392): 1318-21. 349. Du K , Montminy M . C R E B is a regulatory target for the protein kinase Akt/PKB. J Biol Chem 1998; 273(49):32377-9. 350. Pap M , Cooper G M . Role of glycogen synthase kinase-3 in the phosphatidylinositol 3-Kinase/Akt cell survival pathway. J Biol Chem 1998; 273(32): 19929-32. 351. Brunet A , Bonni A , Zigmond M J , Lin M Z , Juo P, et al. Akt promotes cell survival by phosphorylating and inhibiting a Forkhead transcription factor. Cell 1999; 96(6):857-68. 352. Romashkova JA, Makarov SS. NF-kappaB is a target of A K T in anti-apoptotic PDGF signalling. Nature 1999; 401(6748):86-90. 353. Borgatti P, Zauli G, Colamussi M L , Gibellini D, Previati M , et al. Extracellular HIV-1 Tat protein activates phosphatidylinositol 3- and Akt/PKB kinases in CD4+ T lymphoblastoid Jurkat cells. Eur J Immunol 1997; 27(11):2805-11. 354. Meili R, Cron P, Hemmings B A , Ballmer-Hofer K. Protein kinase B/Akt is activated by polyomavirus middle-T antigen via a phosphatidylinositol 3-kinase-dependent mechanism. Oncogene 1998; 16(7):903-7. 355. Pullen N , Thomas G. The modular phosphorylation and activation of p70S6K. FEBS Lett 1997;410(l):78-82. 356. Pullen N , Dennis PB, Andjelkovic M , Dufner A , Kozma SC, et al. Phosphorylation and activation ofp70s6kby PDK1. Science 1998; 279(5351):707-10. 357. Westphal RS, Coffee R L , Jr., Marotta A , Pelech SL, Wadzinski BE . Identification of kinase-phosphatase signaling modules composed of p70 S6 kinase-protein phosphatase 2A (PP2A) and p21-activated kinase-PP2A. J Biol Chem 1999; 274(2):687-92. A P P E N D I X Raw data for Chapter 6 - Effect of age of host animal on in vitro RSV infection. RSV immunostaining on cytospin preparations. Adult Juvenile Mature Intermediate Immature Mature Intermediate Immature 1.3 1.7 1.3 13.0 21.0 20.0 2.6 4.0 1.0 11.0 13.0 17.0 0.7 0.3 1.3 10.0 20.0 24.0 0.7 1.3 0.3 9.0 17.0 23.0 1.7 2.0 2.0 10.0 16.0 19.0 0.7 1.7 1.0 - - -Plaque Assay. Adult Juvenile Mature Intermediate Immature Mature Intermediate Immature 80 88 80 220 1710 2500 6 71 68 300 930 5330 20 112 32 160 950 5330 35 0 27 290 110 3110 67 71 17 290 1170 6500 42 67 67 - - -155 Raw data for Chapter 7 - Effect of sex of host animal on in vitro RSV infection. RSV immunostaining on cytospin preparations. Plaque Assay. Number of syncytia per mi Male Female Mature Immature Mature Immature 10.0 23.3 12.0 25.0 9.7 23.0 13.7 25.3 9.0 21.0 13.0 23.0 11.0 19.7 15.0 27.0 lion A M from male and female guinea pigs Male Female Mature Immature Mature Immature 330 6500 420 7100 400 5700 350 6700 410 8300 380 7000 500 6500 330 8600 IL-6 production by RSV-infected A M subpopulations from male and female guinea pigs Male Female Mature Immature Mature Immature 158.5 124.6 179.0 155.9 395.9 349.0 318.0 300.3 458.9 508.1 407.4 395.3 500.3 416.7 491.0 473.1 IL-8 production by RSV-infected A M subpopulations from male and Male Female Mature Immature Mature Immature 682 952 207 507 617 737 582 767 892 1137 957 1457 727 657 762 722 female guinea pigs. T N F a production by RSV-infected A M subpopulations from male and female guinea pigs Male Female Mature Immature Mature Immature 3.2 23.1 3.8 2.2 5.9 4.3 6.6 6.3 40.7 6.5 11.8 18.6 14.2 18.7 12.6 14.3 Raw data for Chapter 8 - PM10-RSV interaction in guinea pig alveolar macrophage Side scatter measurements by flow cytometry. PM10+RSV RSV+PM10 PMIO RSV N E G 88.0 88.8 88.4 76.7 70.2 88.8 94.9 91.9 86.7 86.8 86.0 87.8 86.9 61.9 64.4 83.0 90.5 84.2 73.4 82.6 RSV immunopositivity as measured by flow cytometry. Proportion of RSV-immunopositive A M . PM10+RSV RSV+PM10 PM10 RSV N E G 3.9 12.02 1.89 5.08 0.18 4.95 11.6 1.59 16.41 0.0 3.69 13.7 1.97 7.6 0.15 5.45 12.9 2.48 12.9 0.0 Plaque Assay. PM10+RSV RSV+PM10 PMIO RSV N E G 363 575 0 4375 0 150 725 0 1488 0 400 1038 0 4750 0 650 1700 0 3525 0 IL-6 production by guinea pig A M subjected to different treatments. PM10+RSV RSV+PM10 PMIO RSV N E G 67.0 10.5 9.1 335.9 3.0 78.0 10.5 0.5 337.7 5.2 62.8 9.8 0.3 355.9 9.9 69.5 14.5 0.7 282.3 7.0 IL-8 production by guinea pig A M subjected to difl PM10+RSV RSV+PM10 PMIO RSV N E G 214.2 187.9 32.8 534.6 288.6 185.4 209.2 66.2 344.6 199.8 217.3 187.3 41.2 470.9 334.6 282.3 292.3 41.2 530.9 201.7 'erent treatments. TNF production by guinea pig A M subjected to difl PM10+RSV RSV+PM10 PMIO RSV N E G 232.0 264.0 249.0 317.8 119.3 308.9 351.3 333.1 247.9 235.4 282.2 372.9 251.7 422.6 150.0 346.3 197.2 281.0 298.4 105.8 'erent treatments. 157 Raw data for Chapter 9 - Summary of A M data from KPKS-1.0 Species GuineaPig GuineaPig GuineaPig Species GuineaPig GuineaPig GuineaPig Tissue Lung Lung Lung Tissue Lung Lung Lung Cell line A M A M A M Cell line A M A M A M KPKS1 gels 368,369 372,373 376,377 KPKS1 gels 368,369 372,373 376,377 KinexusID 330 332 332 KinexusID 330 332 332 Untreated RSV RSV Untreated RSV RSV Exposure 160 sec 160 sec 300 sec Exposure 160 sec 160 sec 300 sec RafB 0 0 0 p52 S6K 7029 500 500 Erk3 0 0 0 Rskl 5365 0 0 p45 0 0 0 Rsk2 0 3780 5918 p43 1559 500 6170 Cot 7192 4941 7203 PKB-alpha 0 1007 1646 Pirn 1 0 0 0 Erkl -A2 2180 2513 4278 CK2-alpha 5344 11603 19759 Erk2-A2 2515 7704 12950 CK2-alpha' 6140 0 0 Erk2-A3 3357 4422 7750 CK2-alpha\" 0 0 0 PDK1 0 0 0 E r k l - B l 3279 1702 2856 Cdkl 0 0 0 Erk2-Bl 3284 4987 8766 Cdk2 6244 0 0 Yes 0 0 0 Cdk4 0 3988 4191 IKK-alpha 0 0 0 Cdk5 4042 0 3580 D A P K 0 0 0 Cdk6 1263 1467 2181 Lck 0 0 0 Cdk7 6082 2874 4822 B M X 1495 2157 3743 Cdk9 0 0 0 Ksr l 0 0 0 p38-alpha 15004 11289 18506 Csk 0 0 0 PKC-mu 0 0 0 Lyn 6815 6831 11611 PRC-alpha 3779 2808 4558 Fyn-L 5431 500 500 PKC-beta 11726 4041 6285 Fvn-U 6610 8608 15820 PKC-gamma 0 0 0 Svk 3894 500 6026 PKC-epsilon 0 0 0 Pyk2 0 0 0 PKC-delta 0 0 0 Btk 0 0 0 PKC-lambda 0 0 0 CK1-delta 0 0 0 PKC-theta 0 0 0 CKI-epsilon 0 0 0 PKC-zeta 3096 2736 3947 ZAP70 0 0 0 Erk6 0 0 0 G C K 0 0 0 SAPKp46 0 1344 0 CaMK4 0 0 0 SAPKp54 0 0 0 C a M K K 0 0 0 Mos 0 0 0 C a M K l 0 0 0 Mek l 1264 4473 7037 Mnk2 0 0 0 Mek2 0 0 0 Nek2 0 0 0 Mek4 0 0 0 Rafl 16108 7197 11971 Mek6 5922 0 0 P K G 0 0 0 Mek7 6 0 0 P K A 1044 1831 2902 PAK-alpha 7599 3781 5698 F A K 500 4567 500 ROK-alpha 0 0 0 GRK2 6219 2998 4833 158 R K R 0 0 0 HDK 5181 0 0 GSK3-alpha 4662 4212 6135 ZIPK 37748 11193 21213 GSK3-beta 0 0 0 JAK1 17068 500 500 D90 S6K 9428 5614 8468 JAK2 0 0 0 D70 S6K U 0 0 0 Src 0 1210 0 p70 S6K L 0 3686 5427 Mstl 0 0 0 "@en . "Thesis/Dissertation"@en . "2002-05"@en . "10.14288/1.0090551"@en . "eng"@en . "Experimental Medicine"@en . "Vancouver : University of British Columbia Library"@en . "University of British Columbia"@en . "For non-commercial purposes only, such as research, private study and education. Additional conditions apply, see Terms of Use https://open.library.ubc.ca/terms_of_use."@en . "Graduate"@en . "In vitro respiratory syncytial virus infection in guinea pig alveolar macrophages"@en . "Text"@en . "http://hdl.handle.net/2429/13081"@en .