Announcement

Collapse
No announcement yet.

Autocatalytic Activation of Influenza Hemagglutinin

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

  • Autocatalytic Activation of Influenza Hemagglutinin

    Journal of Molecular Biology
    Article in Press, Corrected Proof - Note to users
    <!-- TABLE WITH 'THIS DOCUMENT' AND OTHER INFO --><TABLE cellSpacing=0 cellPadding=0 width="100%" valign="bottom"><TBODY><TR><TD>
    </TD></TR></TBODY></TABLE><TABLE cellSpacing=0 cellPadding=0 width=230 align=right border=0 valign="top"><TBODY><TR vAlign=bottom bgColor=#ddf8ad><TD vAlign=bottom align=right width=6 bgColor=#ffffff colSpan=2 height=6 rowSpan=2></TD><TD bgColor=#8cc919 colSpan=2 height=1></TD><TD vAlign=bottom align=left width=6 bgColor=#ffffff colSpan=2 height=6 rowSpan=2></TD></TR><TR bgColor=#ddf8ad><TD colSpan=2 height=5></TD></TR><TR bgColor=#ddf8ad><TD width=1 bgColor=#8cc919></TD><TD width=5></TD><TD class=txtBold noWrap align=left colSpan=2>This Document</TD><TD width=5></TD><TD width=1 bgColor=#8cc919></TD></TR><TR bgColor=#ddf8ad><TD width=1 bgColor=#8cc919></TD><TD colSpan=4 height=3></TD><TD width=1 bgColor=#8cc919></TD></TR><TR><TD bgColor=#8cc919 colSpan=6></TD></TR><TR><TD width=1 bgColor=#8cc919></TD><TD width=5></TD><TD class=txt vAlign=center align=middle width=20></TD><TD class=txt vAlign=center align=left>SummaryPlus </TD><TD width=5></TD><TD width=1 bgColor=#8cc919></TD></TR><TR><TD width=1 bgColor=#8cc919></TD><TD width=5></TD><TD vAlign=center align=middle width=20></TD><TD vAlign=center align=left>Full Text + Links </TD><TD width=5></TD><TD width=1 bgColor=#8cc919></TD></TR><TR><TD width=1 bgColor=#8cc919></TD><TD width=5></TD><TD width=20></TD><TD class=txt> ?Thumbnail Images</TD><TD width=5></TD><TD width=1 bgColor=#8cc919></TD></TR><TR><TD width=1 bgColor=#8cc919></TD><TD width=5></TD><TD vAlign=center align=middle width=20></TD><TD vAlign=center align=left>PDF (622 K) </TD><TD width=5></TD><TD width=1 bgColor=#8cc919></TD></TR><TR><TD width=1 bgColor=#8cc919></TD><TD colSpan=4 height=3></TD><TD width=1 bgColor=#8cc919></TD></TR><TR><TD bgColor=#8cc919 colSpan=6></TD></TR><TR bgColor=#ddf8ad><TD width=1 bgColor=#8cc919></TD><TD colSpan=4 height=3></TD><TD width=1 bgColor=#8cc919></TD></TR><TR bgColor=#ddf8ad><TD width=1 bgColor=#8cc919></TD><TD width=5></TD><TD class=txtBold align=left colSpan=2>Actions</TD><TD width=5></TD><TD width=1 bgColor=#8cc919></TD></TR><TR bgColor=#ddf8ad><TD width=1 bgColor=#8cc919></TD><TD colSpan=4 height=3></TD><TD width=1 bgColor=#8cc919></TD></TR><TR><TD bgColor=#8cc919 colSpan=6></TD></TR><TR><TD width=1 bgColor=#8cc919></TD><TD width=5></TD><TD vAlign=center align=middle></TD><TD class=txt vAlign=center align=left>Cited By</TD><TD width=5></TD><TD width=1 bgColor=#8cc919></TD></TR><TR><TD width=1 bgColor=#8cc919></TD><TD width=5></TD><TD vAlign=center align=middle></TD><TD class=txt vAlign=center align=left>Save as Citation Alert</TD><TD width=5></TD><TD width=1 bgColor=#8cc919></TD></TR><TR vAlign=top><TD width=1 bgColor=#8cc919></TD><TD width=5></TD><TD vAlign=center align=middle></TD><TD class=txt vAlign=center align=left>E-mail Article</TD><TD width=5></TD><TD width=1 bgColor=#8cc919></TD></TR><TR vAlign=top><TD width=1 bgColor=#8cc919></TD><TD width=5></TD><TD vAlign=center align=middle></TD><TD class=txt vAlign=center align=left>Export Citation</TD><TD width=5></TD><TD width=1 bgColor=#8cc919></TD></TR><TR vAlign=top><TD width=1 bgColor=#8cc919></TD><TD width=5></TD><TD vAlign=center align=middle></TD><TD vAlign=center align=left></TD><TD width=5></TD><TD width=1 bgColor=#8cc919></TD></TR><TR vAlign=top><TD vAlign=top align=right width=6 bgColor=#ffffff colSpan=2 height=6 rowSpan=2></TD><TD colSpan=2 height=5></TD><TD vAlign=top align=left width=6 bgColor=#ffffff colSpan=2 height=6 rowSpan=2></TD></TR><TR><TD bgColor=#8cc919 colSpan=2 height=1></TD></TR></TBODY></TABLE><!-- END OF TABLE WITH 'THIS DOCUMENT' AND OTHER INFO -->
    doi:10.1016/j.jmb.2006.09.015

    Copyright ? 2006 Elsevier Ltd All rights reserved.
    Communication



    Autocatalytic Activation of Influenza Hemagglutinin


    Jeong H. Lee<SUP>1</SUP>, Mark Goulian<SUP>2</SUP><SUP>, </SUP><SUP>4</SUP> and Eric T. Boder<SUP>1</SUP><SUP>, </SUP><SUP>3</SUP><SUP>, </SUP><SUP>4</SUP><SUP>, </SUP><SUP></SUP><SUP>, </SUP><SUP></SUP>

    <SUP>1</SUP>Department of Chemical and Biomolecular Engineering, University of Pennsylvania, 220 South 33rd Street, Philadelphia, PA 19104, USA
    <SUP>2</SUP>Department of Physics, University of Pennsylvania, 209 South 33rd Street, Philadelphia, PA 19104, USA
    <SUP>3</SUP>Department of Bioengineering, University of Pennsylvania, 3320 Smith Walk, Philadelphia, PA 19104, USA
    <SUP>4</SUP>Institute for Medicine and Engineering, University of Pennsylvania, 3340 Smith Walk, Philadelphia, PA 19104, USA

    Received 19 January 2006; revised 5 September 2006; accepted 5 September 2006. Edited by I. Wilson. Available online 12 September 2006.



    Abstract


    Enveloped viruses contain surface proteins that mediate fusion between the viral and target cell membranes following an activating stimulus. Acidic pH induces the influenza virus fusion protein hemagglutinin (HA) via irreversible refolding of a trimeric conformational state leading to exposure of hydrophobic fusion peptides on each trimer subunit. Herein, we show that cells expressing fowl plague virus HA demonstrate discrete switching behavior with respect to the HA conformational change. Partially activated states do not exist at the scale of the cell, activation of HA leads to aggregation of cell surface trimers, and newly synthesized HA refold spontaneously in the presence of previously activated HA. These observations imply a feedback mechanism involving self-catalyzed refolding of HA and thus suggest a mechanism similar to the autocatalytic refolding and aggregation of prions.
    Keywords: cooperative activation; HA; hemagglutinin; influenza; autocatalysis Abbreviations: HA, hemagglutinin; FPV, fowl plague virus; IRES, internal ribosome entry site; CMV, cytomegalovirus; mAb, monoclonal antibody



    Article Outline

    <DL><DT>Assay of HA refolding decoupled from membrane fusion <DT>Cell-scale synchronization of HA activation <DT>Density dependence of HA refolding <DT>Activation-induced HA aggregation <DT>Autocatalytic refolding of HA dependent on fusion peptide <DT>Prion-like refolding behavior <DT>Acknowledgements <DT>References</DT></DL>



    Membrane fusion is an important step in many biological activities, and numerous studies have been performed to investigate fusion protein function.<SUP>1.</SUP><SUP>, </SUP><SUP>2.</SUP><SUP>, </SUP><SUP>3.</SUP><SUP> and </SUP><SUP>4.</SUP> Fusion protein activity is required in all infections by enveloped viruses, including influenza viruses. Influenza, an orthomyxovirus, is surrounded by a lipid bilayer membrane envelope captured from the plasma membrane of the cell that generated it, and infection of a host cell involves binding to cell surface sialic acid-containing glycoproteins followed by internalization into endosomal vesicles. Thus, two membranes (viral and endosomal) separate the viral genome from the cell cytoplasm. Viral genetic material is passed through these barriers to the cytosol by fusion of the viral and endosomal membranes, a thermodynamically unfavorable process in the absence of additional energy release.<SUP>5</SUP> Fusion is promoted by the hemagglutinin (HA) protein, which undergoes a dramatic, irreversible conformational rearrangement following exposure to the low pH of the acidified endosome.<SUP>6</SUP> Viral fusion proteins, including HA, contain highly conserved hydrophobic peptide segments of 20 amino acid residues that are essential to fusion activity.<SUP>1</SUP><SUP>,</SUP><SUP>6</SUP> Current models suggest that these fusion peptides spontaneously insert into either the target or viral membrane upon exposure<SUP>3</SUP> and, along with other conformational alterations of the protein, disrupt membrane structure to allow fusion.
    HA is a homotrimer of heterodimer subunits; each HA subunit is synthesized as a single polypeptide (HA0) and proteolytically cleaved into disulfide-linked HA1 and HA2 subunits.<SUP>7.</SUP><SUP>, </SUP><SUP>8.</SUP><SUP> and </SUP><SUP>9.</SUP> The X-ray crystal structures of neutral pH (inactive) and pH-activated HA fragments indicate that the fusion peptide relocates from a site largely buried within the trimer to the membrane-distal end of the structure upon activation.<SUP>6</SUP><SUP>,</SUP><SUP>10</SUP> Experimental evidence indicates this refolding pathway contains reversible intermediate steps,<SUP>10.</SUP><SUP>, </SUP><SUP>11.</SUP><SUP>, </SUP><SUP>12.</SUP><SUP> and </SUP><SUP>13.</SUP> and cooperative interaction of adjacent trimers is believed to promote the irreversible transition to the final fusion-competent, refolded state.<SUP>14.</SUP><SUP>, </SUP><SUP>15.</SUP><SUP>, </SUP><SUP>16.</SUP><SUP>, </SUP><SUP>17.</SUP><SUP> and </SUP><SUP>18.</SUP> These observations are consistent with the putative requirement for synchronized energy release via conformational change of multiple HA trimers to overcome the thermodynamic barrier to membrane fusion.<SUP>5</SUP> Assay of HA refolding decoupled from membrane fusion


    We have developed a mammalian expression system where the conformational change of HA from the H7N1 antigenic subtype fowl plague virus A/FPV/Rostock/34 (FPV), can be detected easily and quantitatively by immunofluorescence. A ten amino acid epitope tag sequence (c-myc) was substituted for the C-terminal half of the 20 amino acid FPV fusion peptide (Figure 1(a) and (b)), thereby generating a presumably non-fusogenic mutant HA (designated HA3.1) with a conveniently detectable tag potentially sensitive to the activation state of the protein. Cell surface HA (both tagged mutant and wild-type) expressed on transiently transfected NIH-3T3 fibroblasts was analyzed by flow cytometry. Immunofluorescent staining with an anti-HA1 monoclonal antibody demonstrated indistinguishable expression levels of the HA3.1 mutant and wild-type HA (data not shown). The mutant HA also showed irreversible, pH-induced exposure of the epitope tag indistinguishable from pH-induced activation of wild-type HA (Figure 1(c)), and pH-activated mutant HA were quantitatively sensitive to thermolysin digestion (0.1 mg/ml), which removes the first 23 residues of the HA2 subunit,<SUP>19</SUP> and reduction of the HA1?HA2 intrachain disulfide bond (data not shown). Non-activated mutant HAs were insensitive to thermolysin or DTT, consistent with the behavior of wild-type HA from various influenza strains.<SUP>17</SUP> Two alternative mutants constructed with the epitope tag in the first 11 fusion peptide residues expressed but failed to demonstrate c-myc tag exposure at any tested pH (data not shown), possibly due to inhibition of proteolytic cleavage of the HA0 form, the necessity of fusion peptide residues 1?11 in the activation, or perturbation of interactions between these residues and the charged pocket of the HA1 subunit.<SUP>20</SUP> Varying the pH of the activating pulse applied to the cells yielded transitions dependent upon the length of the pulse, as expected for an irreversible process (Figure 1(c)).



    Figure 1. Functional analysis of wild-type and HA3.1 mutant hemagglutinin and density dependence of c-myc induction. (a) The c-myc epitope tag sequence (underlined) replaced a ten amino acid sequence of the fusion pepide. A Kozak sequence GCCACC was added to the wild-type FPV HA gene<SUP>34</SUP> (from A/FPV/Rostock/34) by PCR in order to increase the expression in mammalian cells. Amino acid residues 343 to 352, 344 to 353, or 353 to 362 (i.e. fusion peptide residues 1?10, 2?11, or 11?20) were replaced with the c-myc epitope tag sequence via PCR using a SalI site introduced by a silent mutation just upstream of the fusion peptide region. Wild-type HA and the mutants were cloned into a pIRES (BD Biosciences, Mountain View, CA) based mammalian expression vector containing the CMV promoter after the IRES sequence was removed. (b) Structural rendering of X-31 HA (homologous to FPV HA) highlighting the location of the fusion peptide in the pre-activation conformation. Fusion peptide residues 1?10 are colored blue and 11?20 (mutated in HA3.1) are colored red. Coordinates used in the analysis are from PDB file 1HGF,<SUP>35</SUP> and the images were generated using MacPyMOL software (http://www.pymol.org). (c) Cells were transfected using Lipofectamine 2000 (Invitrogen, Carlsbad, CA) following the manufacturer's protocol. For negative control experiments, one well was transfected with empty pIRES vector. Transfection efficiency was monitored by parallel transfection with gWiz-GFP (Gene Therapy Systems, San Diego, CA). Adherent, non-confluent HA3.1 (filled symbols) or wild-type HA (open symbols) vector transient transfectants (24 h post-transfection) were pH-pulsed for 2 min or 5 min at room temperature. HA3.1 cells were double-stained with rabbit polyclonal anti-FPV HA and anti-c-myc mAb 9E10 (Covance, Berkeley, CA), and wild-type HA cells were double-stained with biotinylated anti-HA1 mAb HC2 and conformation-specific anti-HA1 mAb HC58 (specific for trimeric HA1 in the pre-activation conformation).<SUP>36</SUP> HC2 mAb was purified from ascites fluid using Protein G-agarose (Invitrogen, Carlsbad, CA) and biotinylated using the EZ-Link NHS-PEO4-Biotinylation Kit (Pierce, Rockford, IL) following the manufacturer's protocol. Only viable cells, as determined by propidium iodide exclusion and/or light scattering, were included in the analysis by flow cytometry, and 20,000 gated events were analyzed for all flow cytometric experiments. The fraction of HA3.1-expressing cells positive for the c-myc epitope or the fraction of wild-type HA-expressing cells negative for HC58 is shown. Data represent averages of four independent experiments and error bars represent the standard deviation. (d) Wild-type HA-expressing cells (WT HA) were pulsed at the indicated pH for 5 min at room temperature, washed with pH 7.4 buffer, co-stained with biotin-HC2 and HC58, and analyzed by flow cytometry. Fractions of the total viable cell population falling within the indicated regions of the two-color histograms are indicated. (e) HA3.1-expressing cells were pulsed at the indicated pH for 5 min or (f) pulsed at pH 5.0 for the indicated time at room temperature and co-stained with rabbit anti-HA and anti-c-myc. (g) HA3.1-expressing cells were pulsed at the indicated pH for 5 min at room temperature, co-stained with HC58 and chicken anti-c-myc (Jackson ImmunoResearch Laboratory, West Grove, PA), and analyzed by flow cytometry.
    Cell-scale synchronization of HA activation


    Surprisingly, we observed that transfected cells subjected to pH pulses in the transition region (i.e. partially activated states of the system) demonstrate a strict bimodal response at the cell population level in the case of both tagged mutant and wild-type HA (Figure 1(d)?(g)). While fewer cells are activated by higher pH pulses, cell response is rigorously two-state; the intermediate states of the system (i.e. cells activated at intermediate pH or for intermediate duration) include no cells with intermediate levels of c-myc epitope exposure or loss of HC58 reactivity, but rather a mixture of cells with fully non-activated or fully refolded HA. In addition, no cells are positive for both c-myc tag exposure and pre-induction conformation (Figure 1(g)). Thus, no cell exists in a state in which only a fraction of its inducible HA trimers have proceeded through the irreversible conformational change. This striking result indicates that individual cells reach a critical threshold beyond which all activation-competent HA trimers on the cell are irreversibly committed to refolding, implying either cooperative activity on a massive scale or an alternative mechanism such as autocatalysis is involved in the refolding pathway. Density dependence of HA refolding


    Concerted activation of FPV HA proceeding to completion following reversal of the pH pulse implies intertrimer interactions and thus a dependence on trimer cell surface density. To further investigate this phenomenon, we quantitatively analyzed density dependence using the flow cytometric data given in Figure 1(e)?(f). As the acidity or duration of the activating pulse increased, cells with decreasing levels of surface-expressed HA were activated (Figure 1(e)?(f) and Figure 2), consistent with the behavior of HA from other antigenic subtype strains as observed via bulk biochemical assays,<SUP>16</SUP><SUP>,</SUP><SUP>17</SUP><SUP>,</SUP><SUP>21</SUP> which yield data averaged over cell or virus populations. Analysis of activation lag times as a function of surface expression density suggests lag time converges to 2.5 min at high HA density (Figure 2(c)).



    Figure 2. Expression density dependence of HA induction. Flow cytometry data were gated at various levels of anti-HA staining intensity (minimum region width set to maintain at least 100 gated events for all data points) using CellQuest software (Becton-Dickinson, Franklin Lakes, NJ) to allow calculation of (a) the fraction of cells induced as a function of relative HA expression density for pulse durations of 0 min (♦), 3.0 min (*), 3.5 min (▪), 4.0 min () and 4.5 min (●), or (b) the fraction of cells induced as a function of pulse time for HA expression densities at normalized mean fluorescence intensities of 2.28 (♦), 25.85 (●), 80.33 (▪) and 204.43 (). A dose-response equation was fit to the data with two adjustable parameters (pulse time required for 50% of the cell population to be induced [T50] and inflection slope). The minimum R<SUP>2</SUP> value of the fitted curves was 0.9875. (c) T50 values calculated from three separate experiments were averaged and plotted as a function of relative HA expression density. Vertical error bars represent the standard deviation of the three values and horizontal error bars represent the standard deviation of the average corrected mean fluorescence intensity (MFL) of cells within a density gate from three experiments. Absence of error bars indicates range is less than the width of the symbol. The MFL for each gate was corrected by subtracting the mean background auto-fluorescence intensity of a stained negative control population.
    Activation-induced HA aggregation


    The pre and post-activation spatial localization of FPV HA were examined by fluorescence microscopy (Figure 3). Cells expressing either wild-type or c-myc-tagged FPV HA demonstrate distributed anti-HA staining prior to an activating pulse, while acid-pulsed cells demonstrate substantial aggregation of HA. Thus, pH-induced refolding of HA appears to coincide with aggregation of refolded trimers, in agreement with previous observations of pH-induced morphological alteration and mobility restriction of HA from another influenza strain.<SUP>5</SUP><SUP>,</SUP><SUP>19</SUP><SUP>,</SUP><SUP>22</SUP> HA3.1-expressing cells activated at intermediate pH and co-stained for both HA3.1 protein and exposed c-myc tag demonstrate the expected discrete activation phenomenon, and exposed c-myc tags co-localize with aggregated HA3.1 protein (Figure 3(c)). In accord with flow cytometry data, partially activated cell populations show cells positive for either aggregated and exposed c-myc tags or distributed and non-refolded HA3.1, with no coincidence of the two phenotypes (Figure 3(d)).



    Figure 3. Aggregation of HA following pH pulse. (a) Live HA3.1-expressing or wild-type HA-expressing fibroblasts were immunofluorescently stained with anti-c-myc antibody (HA3.1) or anti-HA serum (WT HA) following a 5 min pulse at pH 5.0 and room temperature and analyzed at 20? magnification on a Nikon Eclipse ES300 inverted fluorescence microscope. (b) HA3.1 or wild-type HA-expressing cells were stained with anti-HA serum following a 5 min pulse at the indicated pH and room temperature and analyzed at 64? magnification on an Olympus IX71 inverted fluorescence microscope. (c) HA3.1 expressing cells were co-stained with anti-c-myc/goat anti-mouse-TRITC (red) followed by biotin-HC2/streptavidin-FITC (green) following a 5 min pulse at pH 5.4 (room temperature) and analyzed at 40? magnification on an Olympus IX71 inverted fluorescence microscope. FITC and TRITC fluorescence images were overlayed using NIH ImageJ v1.36 (http://www.rsb.info.nih.gov/ij/). (d) Cells were pulsed at pH 5.4 for 5 min at room temperature 24 h post-transfection, co-stained with biotin-HC58/streptavidin-FITC (BD Biosciences, San Diego, CA) and 9E10/anti-mouse-TRITC (Sigma, St. Louis, MO) , and analyzed as for (c).
    Autocatalytic refolding of HA dependent on fusion peptide


    What mechanism might explain the observed phenomena? Intertrimer cooperativity leading to tandem refolding of trimer aggregates is consistent with density dependence; however, this model predicts that partially induced intermediate states, in which only a fraction of cooperative units on a given cell are refolded, should exist at the cell scale assuming the cooperative unit is smaller than the number of trimers expressed. We observe no such intermediate states in a large statistical sampling by flow cytometry, as indicated by the lack of intermediate cell populations demonstrating c-myc staining de-correlated from expression level (Figure 1(d)?(f)). The absence of partially activated cellular intermediates would thus suggest that the minimal cooperative unit contains all trimers on the cell surface; however, this model then raises the question as to how the cooperative unit could vary dramatically between two cells expressing HA at levels differing by more than tenfold, as seen here (e.g. Figure 1(e)). We observe no stable (i.e. equilibrated) cell populations where high expressers are activated while those of low expression density are not, suggesting the critical cooperative unit size must be less than the minimum observable trimer expression level in the transfected cell population. Our observation that refolding of FPV HA trimers on a cell proceeds irreversibly to a fully activated state, regardless of expression density, therefore argues that cooperative activation cannot fully explain the phenomenon.
    The simplest model consistent with these data includes a positive feedback mechanism by which irreversibly activated FPV HA trimers (which may result from an event involving a small cooperative unit) promote the irreversible refolding of non-activated trimers (Figure 4(a)). This model accounts for the observed irreversible commitment to refolding at the cell scale; once one or more trimers become irreversibly refolded, the feedback chain reaction proceeds until all remaining activatable trimers have refolded, despite removal of the activating pH stimulus.



    Figure 4. Fusion peptide-dependent autocatalytic refolding. (a) A possible refolding reaction pathway for FPV HA is diagrammed. Acidic pH-activated, refolded HA catalyzes refolding of non-activated HA directly or through alteration of lipid bilayer structure and membrane tension. Steps in the shaded region represent putative intermediates suggested in studies of X-31 HA<SUP>12.</SUP><SUP>, </SUP><SUP>13.</SUP><SUP>, </SUP><SUP>14.</SUP><SUP> and </SUP><SUP>15.</SUP> and include exposure of the fusion peptide and clustering into cooperative units (however, reversible intermediates have not been demonstrated for HA from FPV). Cooperative units might include both ?primed? and inactive trimers. Data herein implies a positive feedback loop (i.e. an alternative product-catalyzed reaction pathway) must be active regardless of intermediate states in the conformational change. (b)?(d) HA3.1-transfected cells were analyzed by flow cytometry 16 h (at which time cells express low but detectable levels of HA3.1) or 24 h (high expression) after transfection. Filled grey histograms represent empty vector-transfected cells. (b) Cells pulsed at 16 h post-transfection and re-cultured for an additional 8 h were stained with HC2 mAb and analyzed by flow cytometry to confirm increased expression of HA3.1 following reculturing. (c) Cells were pulsed at pH 5.0 at 16 h post-transfection. A fraction of the cells were immediately labeled with anti-c-myc antibody and analyzed by flow cytometry. Remaining cells were cultured for an additional 8 h in fresh medium prior to antibody staining and analysis with (P: +) or without (P: −) an additional pH 5 pulse at 24 h post-transfection. (d) Cells pH-pulsed at 16 h post-transfection as for (c) were treated with (T: +) or without (T: −) thermolysin (0.1 mg/ml for 30 min at 37 ?C; EMD Biosciences, San Diego, CA) to cleave the fusion peptide, and a fraction of the cells were analyzed by anti-c-myc staining and flow cytometry. Remaining cells were cultured for an additional 8 h prior to staining and analyzed with (P: +) or without (P: −) an additional pH pulse at 24 h post-transfection. (e) Cells transfected with wild-type HA vector were treated as for (c) and (d) and analyzed using mAb HC58 specific for non-refolded HA and mAb HC2 specific for total HA protein to confirm autocatalytic and thermolysin-sensitive refolding of the wild-type protein.
    To more directly test this model, we pulsed cells expressing low levels of HA prior to further culturing to allow synthesis and secretion of additional HA that had not been exposed to the inducing low pH pulse. At both 8 h (Figure 4(b) and (c)) and 20 h (data not shown) after the inducing pH pulse, HA3.1-expressing cells demonstrated increased HA levels and a correlated increase in c-myc tag exposure. Similarly, wild-type HA-expressing cells demonstrated increased HA levels 8 h after an inducing pH pulse, yet the new HA failed to react with trimer-specific mAb HC58 (Figure 4(e)), indicating it had adopted the activated conformation. Thus, new HA expressed during the post-pulse culturing refolded spontaneously, supporting the model that refolding was catalyzed by the previously refolded HA already present on the cell. Furthermore, treatment of pH-pulsed cells with thermolysin prior to re-culturing quantitatively cleaved the fusion peptide/c-myc tag region and abolished spontaneous refolding of newly synthesized HA (both HA3.1 and wild-type HA), while the new HA remained inducible by a subsequent acid pulse (Figure 4(d) and (e)). In contrast, trypsin treatment (20 μg/ml, 15 min at 37 ?C) of pH-pulsed cells (which is expected to cleave the HA1 domain<SUP>19</SUP>) had no effect on spontaneous HA3.1 refolding following an additional 8 h in culture (data not shown). FPV HA was previously shown to refold in the acidic Golgi compartment when expressed in CV-1 cells, and this refolding was inhibited by acidotropic agents such as amantidine and NH<SUB>4</SUB>Cl,<SUP>23</SUP> which buffered the intracellular pH. The formal possibility therefore exists that pH pulsing 3T3 cells might alter the intracellular pH to induce HA refolding several hours later, and that thermolysin might inhibit this effect in some manner. Repeating experiments with 3T3 cells maintained in the presence of 10 mM NH<SUB>4</SUB>Cl at all stages except the pH 5 pulse, however, yielded results identical to those of Figure 4(c) (data not shown), suggesting alteration of the intracellular pH is not responsible for the phenomenon. Taken together, these data suggest that the fusion peptide region, and perhaps its effect on the local membrane environment, is necessary for autocatalytic refolding.
    Autocatalysis of HA refolding could result from either direct trimer?trimer interaction or from effects mediated by locally altered membrane structure and tension. Either mechanism suggests the necessity of physical interaction or proximity between refolded and non-refolded trimers and thus is consistent with the observed activation-correlated aggregation (Figure 3). The relative timing of aggregation versus irreversible refolding, however, has yet to be determined for this system, and thus a direct link between autocatalytic refolding and aggregation has not been established. Nonetheless, it is conceivable that trimer refolding could be linked to incorporation into a central protein aggregate or an expanding ring of perturbed membrane structure. The speculation that altered membrane tension might induce refolding is consistent with the membrane destabilization required to induce fusion with a juxtaposed target membrane and has previously been proposed.<SUP>17</SUP> This mechanism would coordinate conformational energy release of a large number of HA trimers and lead to a domain of high-energy membrane structure. Indeed, a similar mechanism involved in HA-mediated membrane fusion has been suggested,<SUP>5</SUP> and Leikina et al., recently demonstrated that HA trimers from outside the intercellular contact zone are necessary for fusion pore expansion,<SUP>24</SUP> consistent with a requirement for activated HA aggregation. Insertion of exposed fusion peptides into the cell membrane (or N-terminal 11 residues of the fusion peptide in the HA3.1 mutant) could alter membrane tension and curvature with resulting effects on trimer conformational stability; membrane tension has been observed to impact folded stability of bacterial outer membrane protein A.<SUP>25</SUP> This supposition is also consistent with the suggested structure of membrane-embedded HA fusion peptide,<SUP>26.</SUP><SUP>, </SUP><SUP>27.</SUP><SUP>, </SUP><SUP>28.</SUP><SUP> and </SUP><SUP>29.</SUP> in which the N-terminal portion of the fusion peptide embeds into the non-polar membrane core and disrupts lipid structure. However, despite the observed thermolysin sensitivity of the autocatalytic refolding, direct evidence for membrane involvement in this phenomenon and for the existence of autocatalytic HA refolding in the context of intermembrane fusion requires further studies. Prion-like refolding behavior

    Our proposed model for refolding of FPV HA to the fusion-active conformation parallels the hypothesized refolding feedback mechanism by which infectious proteins (prions) induce irreversible refolding and aggregation of structural isoforms,<SUP>30</SUP> and the fusion peptide of HA demonstrates sequence similarity to amyloid β peptide and a prion sequence.<SUP>31.</SUP><SUP>, </SUP><SUP>32.</SUP><SUP> and </SUP><SUP>33.</SUP> To our knowledge, this is the first example of data consistent with such a mechanism for conformational regulation putatively involved in the function of a membrane protein. Whether this phenomenon applies to other viral fusion proteins, or to other membrane proteins functionally dependent upon conformational changes, remains to be investigated. Nonetheless, the existence of autocatalytic, pH-induced refolding of FPV hemagglutinin raises interesting possibilities regarding the understanding of conformational molecular switches.



    Acknowledgements


    We thank Dr Hans-Dieter Klenk for providing the anti-FPV HA serum, Dr Steve Wharton for anti-FPV HA mAbs HC2 and HC58, and Dr Paula Cannon for the cloned FPV HA gene. Funding was provided by NSF NIRT 02-10777.



    References


    1. D.M. Eckert and P.S. Kim, Mechanisms of viral membrane fusion and Its inhibition, Annu. Rev. Biochem. 70 (2001), pp. 777?810. Abstract-EMBASE | Abstract-MEDLINE | Abstract-Elsevier BIOBASE | Full Text via CrossRef | Abstract + References in Scopus | Cited By in Scopus
    2. R. Blumenthal, M.J. Clague, S.R. Durell and R.M. Epand, Membrane fusion, Chem. Rev. 103 (2003), pp. 53?69. Abstract-Compendex | Abstract-MEDLINE | Abstract + References in Scopus | Cited By in Scopus
    3. R. Jahn, T. Lang and T.C. Sudhof, Membrane fusion, Cell 112 (2003), pp. 519?533. SummaryPlus | Full Text + Links | PDF (587 K) | Abstract + References in Scopus | Cited By in Scopus
    4. E.H. Chen and E.N. Olson, Unveiling the mechanisms of cell-cell fusion, Science 308 (2005), pp. 369?373. Abstract-MEDLINE | Abstract-Compendex | Abstract-EMBASE | Abstract-Elsevier BIOBASE | Abstract-GEOBASE | Full Text via CrossRef | Abstract + References in Scopus | Cited By in Scopus
    5. L.V. Chernomordik and M.M. Kozlov, Protein-lipid interplay in fusion and fission of biological membranes, Annu. Rev. Biochem. 72 (2003), pp. 175?207. Abstract-MEDLINE | Abstract-Elsevier BIOBASE | Abstract-EMBASE | Full Text via CrossRef | Abstract + References in Scopus | Cited By in Scopus
    6. J.J. Skehel and D.C. Wiley, Receptor binding and membrane fusion in virus entry: the influenza hemagglutinin, Annu. Rev. Biochem. 69 (2000), pp. 531?569. Abstract-MEDLINE | Abstract-EMBASE | Abstract-Elsevier BIOBASE | Full Text via CrossRef | Abstract + References in Scopus | Cited By in Scopus
    7. R.T. Huang, R. Rott and H.D. Klenk, Influenza viruses cause hemolysis and fusion of cells, Virology 110 (1981), pp. 243?247. Abstract | Abstract + References in Scopus | Cited By in Scopus
    8. R.T. Huang, K. Wahn, H.D. Klenk and R. Rott, Fusion between cell membrane and liposomes containing the glycoproteins of influenza virus, Virology 104 (1980), pp. 294?302. Abstract | Abstract + References in Scopus | Cited By in Scopus
    9. H.D. Klenk, R. Rott, M. Orlich and J. Blodorn, Activation of influenza A viruses by trypsin treatment, Virology 68 (1975), pp. 426?439. Abstract | Abstract + References in Scopus | Cited By in Scopus
    10. J.M. White and I.A. Wilson, Anti-peptide antibodies detect steps in a protein conformational change: low-pH activation of the influenza virus hemagglutinin, J. Cell Biol. 105 (1987), pp. 2887?2896. Abstract-MEDLINE | Abstract-EMBASE | Full Text via CrossRef | Abstract + References in Scopus | Cited By in Scopus
    11. T. Korte, K. Ludwig, F.P. Booy, R. Blumenthal and A. Herrmann, Conformational intermediates and fusion activity of influenza virus hemagglutinin, J. Virol. 73 (1999), pp. 4567?4574. Abstract-Elsevier BIOBASE | Abstract-EMBASE | Abstract-MEDLINE | Abstract + References in Scopus | Cited By in Scopus
    12. T. Korte, K. Ludwig, M. Krumbiegel, D. Zirwer, G. Damaschun and A. Herrmann, Transient changes of the conformation of hemagglutinin of influenza virus at low pH detected by time-resolved circular dichroism spectroscopy, J. Biol .Chem. 272 (1997), pp. 9764?9770. Abstract-MEDLINE | Abstract-EMBASE | Abstract + References in Scopus | Cited By in Scopus
    13. S.A. Tatulian and L.K. Tamm, Reversible pH-dependent conformational change of reconstituted influenza hemagglutinin, J. Mol. Biol. 260 (1996), pp. 312?316. Abstract | PDF (224 K) | Abstract + References in Scopus | Cited By in Scopus
    14. M. Krumbiegel, A. Herrmann and R. Blumenthal, Kinetics of the low pH-induced conformational changes and fusogenic activity of influenza hemagglutinin, Biophys. J. 67 (1994), pp. 2355?2360. Abstract-INSPEC | Abstract-EMBASE | Abstract-MEDLINE | Abstract-Elsevier BIOBASE | Abstract + References in Scopus | Cited By in Scopus
    15. E. Leikina, C. Ramos, I. Markovic, J. Zimmerberg and L.V. Chernomordik, Reversible stages of the low-pH-triggered conformational change in influenza virus hemagglutinin, EMBO J. 21 (2002), pp. 5701?5710. Abstract-EMBASE | Abstract-MEDLINE | Abstract-Elsevier BIOBASE | Full Text via CrossRef | Abstract + References in Scopus | Cited By in Scopus
    16. T. Danieli, S.L. Pelletier, Y.I. Henis and J.M. White, Membrane fusion mediated by the influenza virus hemagglutinin requires the concerted action of at least three hemagglutinin trimers, J. Cell Biol. 133 (1996), pp. 559?569. Abstract-MEDLINE | Abstract-Elsevier BIOBASE | Abstract-EMBASE | Full Text via CrossRef
    17. I. Markovic, E. Leikina, M. Zhukovsky, J. Zimmerberg and L.V. Chernomordik, Synchronized activation and refolding of influenza hemagglutinin in multimeric fusion machines, J Cell Biol 155 (2001), pp. 833?844. Abstract-MEDLINE | Full Text via CrossRef
    18. J. Bentz, Minimal aggregate size and minimal fusion unit for the first fusion pore of influenza hemagglutinin-mediated membrane fusion, Biophys. J. 78 (2000), pp. 227?245. Abstract-MEDLINE | Abstract-EMBASE | Abstract-Elsevier BIOBASE | Abstract + References in Scopus | Cited By in Scopus
    19. R.W. Ruigrok, N.G. Wrigley, L.J. Calder, S. Cusack, S.A. Wharton, E.B. Brown and J.J. Skehel, Electron microscopy of the low pH structure of influenza virus haemagglutinin, EMBO J. 5 (1986), pp. 41?49. Abstract-MEDLINE | Abstract + References in Scopus | Cited By in Scopus
    20. D.A. Steinhauer, Role of hemagglutinin cleavage for the pathogenicity of influenza virus, Virology 258 (1999), pp. 1?20. Abstract | Abstract + References | PDF (489 K) | Abstract + References in Scopus | Cited By in Scopus
    21. M.J. Clague, C. Schoch and R. Blumenthal, Delay time for influenza virus hemagglutinin-induced membrane fusion depends on hemagglutinin surface density, J. Virol. 65 (1991), pp. 2402?2407. Abstract-MEDLINE | Abstract-EMBASE | Abstract + References in Scopus | Cited By in Scopus
    22. O. Gutman, T. Danieli, J.M. White and Y.I. Henis, Effects of exposure to low pH on the lateral mobility of influenza hemagglutinin expressed at the cell surface: correlation between mobility inhibition and inactivation, Biochemistry 32 (1993), pp. 101?106. Abstract-EMBASE | Abstract-MEDLINE | Full Text via CrossRef | Abstract + References in Scopus | Cited By in Scopus
    23. M. Ohuchi, A. Cramer, M. Vey, R. Ohuchi, W. Garten and H.D. Klenk, Rescue of vector-expressed fowl plague virus hemagglutinin in biologically active form by acidotropic agents and coexpressed M2 protein, J. Virol. 68 (1994), pp. 920?926. Abstract-EMBASE | Abstract-MEDLINE | Abstract-Elsevier BIOBASE | Abstract + References in Scopus | Cited By in Scopus
    24. E. Leikina, A. Mittal, M.S. Cho, K. Melikov, M.M. Kozlov and L.V. Chernomordik, Influenza hemagglutinins outside of the contact zone are necessary for fusion pore expansion, J. Biol. Chem. 279 (2004), pp. 26526?26532. Abstract-MEDLINE | Abstract-EMBASE | Abstract-Compendex | Abstract-Elsevier BIOBASE | Full Text via CrossRef | Abstract + References in Scopus | Cited By in Scopus
    25. H. Hong and L.K. Tamm, Elastic coupling of integral membrane protein stability to lipid bilayer forces, Proc. Natl Acad. Sci. USA 101 (2004), pp. 4065?4070. Abstract-Elsevier BIOBASE | Abstract-MEDLINE | Abstract-EMBASE | Full Text via CrossRef | Abstract + References in Scopus | Cited By in Scopus
    26. X. Han, J.H. Bushweller, D.S. Cafiso and L.K. Tamm, Membrane structure and fusion-triggering conformational change of the fusion domain from influenza hemagglutinin, Nature Struct. Biol. 8 (2001), pp. 715?720. Abstract-EMBASE | Abstract-Elsevier BIOBASE | Abstract-MEDLINE | Full Text via CrossRef | Abstract + References in Scopus | Cited By in Scopus
    27. P. Lague, B. Roux and R.W. Pastor, Molecular dynamics simulations of the influenza hemagglutinin fusion Peptide in micelles and bilayers: conformational analysis of Peptide and lipids, J. Mol. Biol. 354 (2005), pp. 1129?1141. SummaryPlus | Full Text + Links | PDF (520 K) | Abstract + References in Scopus | Cited By in Scopus
    28. L. Vaccaro, K.J. Cross, J. Kleinjung, S.K. Straus, D.J. Thomas and S.A. Wharton et al., Plasticity of influenza haemagglutinin fusion peptides and their interaction with lipid bilayers, Biophys. J. 88 (2005), pp. 25?36. Abstract-EMBASE | Abstract-INSPEC | Abstract-MEDLINE | Abstract-Elsevier BIOBASE | Abstract + References in Scopus | Cited By in Scopus
    29. Q. Huang, C.L. Chen and A. Herrmann, Bilayer conformation of fusion peptide of influenza virus hemagglutinin: a molecular dynamics simulation study, Biophys. J. 87 (2004), pp. 14?22. Abstract-Elsevier BIOBASE | Abstract-EMBASE | Abstract-MEDLINE | Full Text via CrossRef | Abstract + References in Scopus | Cited By in Scopus
    30. C. Weissmann, Birth of a prion: spontaneous generation revisited, Cell 122 (2005), pp. 165?168. SummaryPlus | Full Text + Links | PDF (93 K) | Abstract + References in Scopus | Cited By in Scopus
    31. G. Forloni, N. Angeretti, R. Chiesa, E. Monzani, M. Salmona, O. Bugiani and F. Tagliavini, Neurotoxicity of a prion protein fragment, Nature 362 (1993), pp. 543?546. Abstract-MEDLINE | Abstract-EMBASE | Full Text via CrossRef | Abstract + References in Scopus | Cited By in Scopus
    32. V.D. Del Angel, F. Dupuis, J.P. Mornon and I. Callebaut, Viral fusion peptides and identification of membrane-interacting segments, Biochem. Biophys. Res. Commun. 293 (2002), pp. 1153?1160. Abstract-EMBASE | Abstract-Elsevier BIOBASE | Abstract-MEDLINE | Abstract + References in Scopus | Cited By in Scopus
    33. O. Crescenzi, S. Tomaselli, R. Guerrini, S. Salvadori, A.M. D'Ursi, P.A. Temussi and D. Picone, Solution structure of the Alzheimer amyloid beta-peptide (1?42) in an apolar microenvironment. Similarity with a virus fusion domain, Eur. J. Biochem. 269 (2002), pp. 5642?5648. Abstract-EMBASE | Abstract-MEDLINE | Abstract-Elsevier BIOBASE | Full Text via CrossRef | Abstract + References in Scopus | Cited By in Scopus
    34. A.H. Lin, N. Kasahara, W. Wu, R. Stripecke, C.L. Empig, W.F. Anderson and P.M. Cannon, Receptor-specific targeting mediated by the coexpression of a targeted murine leukemia virus envelope protein and a binding-defective influenza hemagglutinin protein, Hum. Gene Ther. 12 (2001), pp. 323?332. Abstract-MEDLINE | Abstract-Elsevier BIOBASE | Abstract-EMBASE | Abstract + References in Scopus | Cited By in Scopus
    35. N.K. Sauter, J.E. Hanson, G.D. Glick, J.H. Brown, R.L. Crowther and S.J. Park et al., Binding of influenza virus hemagglutinin to analogs of its cell-surface receptor, sialic acid: analysis by proton nuclear magnetic resonance spectroscopy and X-ray crystallography, Biochemistry 31 (1992), pp. 9609?9621. Abstract-MEDLINE | Abstract-EMBASE | Full Text via CrossRef | Abstract + References in Scopus | Cited By in Scopus 36. R.J. Sugrue, G. Bahadur, M.C. Zambon, M. Hall-Smith, A.R. Douglas and A.J. Hay, Specific structural alteration of the influenza haemagglutinin by amantadine, EMBO J. 9 (1990), pp. 3469?3476. Abstract-MEDLINE | Abstract-EMBASE | Abstract + References in Scopus | Cited By in Scopus




    <SUP></SUP>Corresponding author. Department of Chemical and Biomolecular Engineering, University of Pennsylvania, 220 South 33rd Street, Philadelphia, PA 19104, USA.

  • #2
    Re: Autocatalytic Activation of Influenza Hemagglutinin

    What are the implications here?....in ordinary English?

    When they say "refoldng" does this relate to HA cleavage?

    Might the acidity (or lack of) relate to who does or does not get BF?

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

    Comment


    • #3
      Re: Autocatalytic Activation of Influenza Hemagglutinin

      thank you,Mingus

      a question ( ou remarque )
      the 1918 HA didn't have a clivage site ( tautenberger ) and the virus was bad enough to kill million of people
      would it be worsest with H5n1 ? or should we consider the constellation of 8 genes

      Comment

      Working...
      X