Varela_cellular_visualization_2014.doc

advertisement
1
Cellular Visualization of Macrophage Pyroptosis
2
and IL1 Release in a Viral Hemorrhagic
3
Infection in Zebrafish Larvae
4
Mónica Varela1, Alejandro Romero1, Sonia Dios1, Michiel van der Vaart2, Antonio
5
Figueras1, Annemarie H. Meijer2, Beatriz Novoa1#
6
1 Marine Research Institute, CSIC, Vigo, Spain, 2 Institute of Biology, Leiden University, Leiden, The
7
Netherlands
8
9
10
Corresponding Author: beatriznovoa@iim.csic.es
11
Running Tittle: Macrophage Pyroptosis in Zebrafish Larvae
12
Keywords: Hemorrhagic Virus, Zebrafish, Macrophages, Pyroptosis, IL1 Release, Antiviral Response,
13
RNA Virus, Inflammatory Response, SVCV
14
Abstract word count: 284
15
Text word count: 5097
16
1
17
ABSTRACT
18
Hemorrhagic viral diseases are distributed worldwide with important pathogens,
19
such as Dengue virus or Hantaviruses. The lack of adequate in vivo infection models has
20
limited the research on viral pathogenesis and the current understanding of the
21
underlying infection mechanisms. Although hemorrhages have been associated with the
22
infection of endothelial cells, other cellular types could be the main targets for
23
hemorrhagic viruses. Our objective was to take advantage of the use of zebrafish larvae
24
in the study of viral hemorrhagic diseases, focusing on the interaction between viruses
25
and host cells. Cellular processes, such as transendothelial migration of leukocytes,
26
viral-induced pyroptosis of macrophages and Il1 release, could be observed in
27
individual cells, providing a deeper knowledge of the immune mechanisms implicated
28
in the disease. Furthermore, the application of these techniques to other pathogens will
29
improve the current knowledge of host-pathogen interactions and increase the potential
30
for the discovery of new therapeutic targets.
31
32
IMPORTANCE
33
Pathogenic mechanisms of hemorrhagic viruses are diverse and most of the
34
research regarding interactions between viruses and host cells has been performed in
35
cell lines that might not be major targets during natural infections. Thus, viral
36
pathogenesis research has been limited because of the lack of adequate in vivo infection
37
models. The understanding of the relative pathogenic roles of the viral agent and the
38
host response to the infection is crucial. This will be facilitated by the establishment of
39
in vivo infection models using organisms as zebrafish, which allows the study of the
40
diseases in the context of complete individual. The use of this animal model with other
2
41
pathogens could improve the current knowledge on host-pathogen interactions and
42
increase the potential for the discovery of new therapeutic targets against diverse viral
43
diseases.
44
3
45
INTRODUCTION
46
Endothelial cells and leukocytes are at the front line of defense against
47
pathogens. Most infectious pathogens have some relationship with the endothelium,
48
although not all pathogens are true endothelial invading organisms (1). Infections can
49
produce large changes in endothelial cells function, particularly in relation to
50
inflammation. Inflammatory stimuli can activate endothelial cells, which secrete
51
cytokines and chemokines. Moreover, the importance of endothelial cells in the
52
regulation of leukocyte transmigration to the site of inflammation has given these cells a
53
major role in immunity (2).
54
Host-pathogen interactions are essential for the modulation of immunity (3). The
55
presence of a pathogen alters the immune balance prevailing in a host under normal
56
conditions causing the appearance of diseases, which lead to the death of the host if it is
57
not able to control the infection. Clear examples of these effects are evident in viral
58
hemorrhagic fevers, in which an understanding of the relative pathogenic roles of the
59
viral agent and the host response to the infection is crucial (4). Hemorrhagic
60
manifestations are characteristic of these fevers, which have historically been associated
61
with endothelium infections. Currently, many pathogens causing hemorrhagic diseases
62
do not infect endothelial cells as the primary target (1). Indeed, macrophages and
63
dendritic cells are targets for some hemorrhagic viruses, such as Ebola Virus (5) or
64
Dengue Virus (6, 7).
65
Most of the research regarding interactions between viruses and host cells has
66
been performed in cell lines that might not be major targets during natural infections
67
(8), making it difficult to characterize the behavior of pathogens in individuals and
68
reducing the likelihood of discovering an effective therapeutic target. The lack of a good
4
69
animal model in which the infection can be followed from the beginning and in the
70
context of the whole organism has contributed to these difficulties. Zebrafish (Danio
71
rerio) provide advantages, as these organisms exhibit rapid development, transparency
72
and a high homology between its genome and the human one (9). Furthermore, the
73
existence of mutant and transgenic fish lines and the potential use of reverse genetics
74
techniques, such as the injection of antisense morpholino oligonucleotides or synthetic
75
mRNA, make zebrafish a useful tool for the study of host-pathogen interactions.
76
No naturally occurring viral infections have been characterized for zebrafish
77
(10). Moreover, the experimental susceptibility of zebrafish to viral infections has been
78
demonstrated with other fish viruses (11-13) and also with mammalian viruses (14-16).
79
We used the rhabdovirus Spring Viraemia of Carp Virus (SVCV) as a viral
80
model to show the potential of the zebrafish in host-pathogen interaction studies. SVCV
81
predominantly affects cyprinid fish and causes important economical losses and
82
mortalities worldwide (17, 18). Affected fish exhibit destruction of tissues, such as
83
kidney or spleen, leading to hemorrhage and death. However, little is known about the
84
infection mechanism of this enveloped RNA virus.
85
Thus, the main objective of the present study was to improve the knowledge of
86
the pathology generated by SVCV, as an example of hemorrhagic virus, focusing on
87
interactions between the virus and host cells at early stages of the infection in zebrafish
88
larvae. Taking advantage of zebrafish transgenic lines, imaging techniques and, using
89
morpholino-induced cell depletion, viral-induced pyroptosis and IL1β release were
90
observed in macrophages at the single cell level.
91
5
92
MATERIALS AND METHODS
93
Ethics Statement. The protocols for fish care and the challenge experiments were
94
reviewed and approved by the CSIC National Committee on Bioethics under approval
95
number 07_09032012. Experiments were conducted in fish larvae before independently
96
feeding and therefore, before an ethical approval is required (EU directive 2010_63).
97
Fish. Homozygous embryos and larvae from wild-type zebrafish, transgenic line
98
Tg(Fli1a:eGFP)y1 that labels endothelial cells, transgenic line Tg(Mpx:GFP)i114 that
99
labels neutrophils, transgenic line Tg(Mpeg1:eGFP)gl22 that labels macrophages and
100
transgenic line Tg(Cd41:GFP)la2 that labels thrombocytes were obtained from our
101
experimental facility, where the zebrafish were cultured using established protocols (19,
102
20). The eggs were obtained according to protocols described in The Zebrafish Book
103
(19) and maintained at 28.5˚C in egg water (5 mM NaCl, 0.17 mM KCl, 0.33 mM
104
CaCl2, 0.33 mM MgSO4, and 0.00005% methylene blue), and 0.2 mM N-
105
phenylthiourea (PTU; Sigma) was used to prevent pigment formation from 1 days post-
106
fertilization (dpf).
107
Virus and Infection. The SVCV isolate 56/70 was propagated on epithelioma
108
papulosum cyprini (EPC) carp cells (ATCC CRL-2872) and titrated in 96-well plates.
109
The plates were incubated at 15°C for 7 days and examined for cytophathic effects each
110
day. After the observation of the cells under the microscope, the virus dilution causing
111
infection of 50% of the inoculated cell line (TCID50) was determined using the Reed-
112
Müench method (21). The fish larvae were infected through microinjection into the duct
113
of Cuvier as described in Benard et al. (22) at 2 or 3 dpf. SVCV was diluted to the
114
appropriate concentration in PBS with 0.1% phenol red (as a visible marker to the
115
injection of the solution into the embryo) just before the microinjection of 2 nl of viral
6
116
suspension per larvae. The infections were conducted at 23˚C. SVCV was heat-killed at
117
65ºC during 20 minutes.
118
Imaging. Images of the signs and videos of the blood flow were obtained using an
119
AZ100 microscope (Nikon) coupled to a DS-Fi1 digital camera (Nikon). Confocal
120
images of live or fixed larvae were captured using a TSC SPE confocal microscope
121
(Leica). The images were processed using the LAS-AF (Leica) and ImageJ software.
122
The 3D reconstructions and the volume clipping were performed using Image Surfer
123
(http://imagesurfer.cs.unc.edu/) and LAS-AF (Leica), respectively.
124
Immunohistochemistry. Whole-mount immunohistochemistry was performed as
125
previously described (23). The following primary antibodies and dilutions were used: L-
126
plastin (rabbit anti-zebrafish, 1:500, kindly provided by Dr. Paul Martin); Caspase a
127
(rabbit anti-zebrafish, 1:500; Anaspec); SVCV (mouse anti-SVCV, 1:500; BioX
128
Diagnostics) and Il1β (rabbit anti-zebrafish, 1:500, kindly provided by Dr. John
129
Hansen). The secondary antibodies used were Alexa 488 anti-rabbit, Alexa 546 anti-
130
rabbit or Alexa 635 anti-mouse (Life Technologies), all diluted 1:1000.
131
Real-Time PCR. Total RNA was isolated from snap-frozen larvae using the Maxwell®
132
16 LEV simply RNA Tissue kit (Promega) with the automated Maxwell® 16
133
Instrument according to the manufacturer’s instructions. cDNA synthesis using random
134
primers and qPCR were performed as previously described (24). 18S was used as a
135
housekeeping to normalize the expression values. Gene expression ratio was calculated
136
by dividing the normalized expression values of infected larvae by the normalized
137
expression values of the controls. Table S1 in supplemental material shows the
138
sequences of the primer pairs used. Two independent experiments, of 3 biological
139
replicates each, were performed.
7
140
Microangiography. Tetramethylrhodamine dextran (2x106 MW, Life Technologies)
141
was injected into the caudal vein of anesthetized zebrafish Tg(Fli1a-GFP) embryos at
142
24 hours post-infection (hpi). The images were acquired within 30 min after injection.
143
TUNEL assay. The TUNEL assay was performed using the In Situ Cell Death
144
Detection Kit, TMR red or POD (Roche) as previously described (25).
145
In vivo acridine orange staining. For acridine orange staining, larvae were incubated
146
for 2 hours in 10 µg/ml acridine orange solution followed by washing for 30 min.
147
Macrophages depletion with PU.1-MO. For the PU.1 knockdown, the morpholino
148
was injected into the yolk at the one-cell stage at 1 mM to block macrophages
149
development as previously described (26). Macrophage-depleted fish were infected at 2
150
dpf to ensure morpholino effectiveness until the end of the infection.
151
In vivo propidium iodide staining. For propidium iodide (PI) staining, larvae were
152
incubated for 2 hours in 10 µg/ml propidium iodide solution followed by washing for 30
153
min.
154
Statistical analysis. Kaplan-Meier survival curves were analyzed with a Log-rank
155
(Mantel Cox) test. The neutrophil count and qPCR data were analyzed using one-way
156
analysis of variance and Tukey’s test, and differences were considered significant at
157
p<0.05. The results are presented as the means ± standard error of mean (SEM).
158
8
159
RESULTS
160
Course of the SVCV infection in zebrafish larvae. Zebrafish larvae are susceptible to
161
SVCV through bath infection, but typically the results show high interindividual
162
variations (12, 27). Moreover, the long infection times required exclude the use of
163
temporal knockdown techniques, such as morpholino microinjection. As with caudal
164
vein injection (27), we found that SVCV injection into the duct of Cuvier was efficient,
165
thereby providing a reproducible route of infection with fast kinetics. We injected 2 nl
166
of a SVCV stock solution (3x106 TCID50/ml) per larva. Mortality began at 22-24 hpi,
167
consistently reaching nearly 100% between 48 and 50 hpi (Fig. 1A). There were no
168
differences in the development of the infection between the use of 2 or 3 dpf larvae
169
(data not shown). To assess viral transcription, the expression of the SVCV
170
nucleoprotein (N gene) was measured through quantitative RT-PCR (Fig. 1B). The
171
samples were collected between 3 and 24 hpi (just prior to mortality). Efficient viral
172
transcription was observed from 6 hpi, with an important increment between 12 and 18
173
hpi. Moreover, we found that viral N gene transcription began quickly in the fish as
174
significant differences were found between heat-killed SVCV (HK-SVCV) and SVCV
175
infected fish (Fig. 1B).
176
With regard to macroscopic signs, the first ones were visible after 18 hours (Fig.
177
1C). A progressive decrease in the blood flow rate was observed, particularly in the tail
178
(See movie S1-S2 in supplemental material). The circulation in the tail was completely
179
stopped after 20-22 hpi, although it remained visible at the anterior part of the fish. The
180
circulation in the head lasted longer but also eventually stopped (See movie S3 in
181
supplemental material). At this point, blood was clearly visible inside the veins, with
182
exception of the hemorrhage areas (Fig. 1C). Hemorrhages appeared in the anterior part
9
183
of the larvae, mostly in areas with a huge amount of microvessels but, a great
184
heterogeneity was observed in the occurrence of these hemorrhages varying in terms of
185
location and time post-infection. The heartbeat was clearly less powerful over time until
186
it completely stopped, moment in which the larvae were considered dead. The observed
187
signs and times are consistent with those previously described for this virus using a
188
similar infection model (27). Consistency in the infection results reaffirms the idea that
189
SVCV and zebrafish larvae are an appropriate duo for an in-depth study of the innate
190
immune response against that intracellular pathogen.
191
SVCV does not affect the integrity of the endothelium. The loss of circulation in the
192
trunk and hemorrhages in larvae head were the most characteristic symptoms of
193
infection. Therefore, we characterized the effect of the virus in the vascular system of
194
the fish, as previously determined for another rhabdovirus, IHNV (13). Using the
195
transgenic line Tg(Fli1a-GFP) (28), which facilitates the visualization of the vascular
196
system, we clearly observed a loss of GFP fluorescence in infected fish at 24 hpi (Fig.
197
1D, E). This effect was particularly visible in the main veins of the head (Fig. 1D) and
198
in the tail (Figure 1E). In addition, we used this transgenic line in combination with a
199
whole-mount fluorescent immunohistochemistry for L-plastin (leukocyte-plastin) (29).
200
The anti-L-plastin antibody was a useful tool for the detection of macrophages and
201
neutrophils in the larvae. In addition to the loss of Fli1a-GFP fluorescence in the trunk,
202
a decrease in the number of L-plastin-positive cells in the area corresponding to the
203
caudal hematopoietic tissue was also observed at 24 hpi (Fig. 1E). Moreover, a shape
204
change in L-plastin-positive cells was clearly visible in infected fish. The morphology
205
of the cells in infected fish changed from thin and elongated (with prolonged
206
extensions) to rounded and lobulated (Fig. 1F). The morphological changes might be
10
207
associated with alterations in cell motility or the infection of the leukocytes themselves
208
(30, 31).
209
To confirm the integrity of the endothelium in the tail at 24 hpi, we injected
210
tetramethylrhodamine dextran into the caudal vein of the larvae. This fluorescent dye
211
remained inside the vessels, even when Fli1a-GFP fluorescence was absent (Fig. 2A).
212
Shadows in tetramethylrhodamine dextran fluorescence correspond with blood cells that
213
remain inside vessels even when circulation was stopped (Fig. 2A, 2B). Moreover, the
214
results of the TUNEL assay and the acridine orange staining revealed no DNA damage
215
neither necrosis, respectively, in the endothelial cells of infected fish (Fig. 2C),
216
suggesting that the cells were apparently healthy and the loss of fluorescence likely
217
reflected a transcriptional regulation of the Fli1a gene and not cell death. This was
218
verified by combining Fli1a-GFP fish with DIC microscopy (Fig. 2D). Further,
219
TUNEL-positive cells in the caudal hematopoietic tissue of infected larvae were
220
observed, and the localization of these cells was consistent with the position of L-
221
plastin-positive hematopoietic precursor cells in control fish (Fig. 2E). This observation
222
was not surprising as blood flow is essential for hematopoietic stem cell development
223
(32). In addition, several studies have shown the relationship between blood circulation
224
and Fli1, as Fli1-deficient mice exhibit impaired hematopoiesis (33). Indeed, previous
225
studies in mice described that both morphants and knockouts of Fli1 showed normal
226
vessel development, with impaired circulation and hemorrhaging (33, 34). This might
227
suggest that the simple transcriptional regulation of fli1a could be sufficient to cause the
228
symptoms observed in SVCV-infected larvae.
229
To determine the mechanism underlying fli1a mRNA transcription and to
230
associate it with the loss of blood flow in the larvae, we performed a time course during
11
231
SVCV infection. Based on the signs of their morphants and on their implication in the
232
maintenance of vessel integrity, we also selected the v-ets erythroblastosis virus E26
233
oncogene homolog 1a (ets1a) and the vascular endothelial cadherin (ve-cad) genes for
234
qPCR analysis (Fig. 2F). Fli1a and Ets1a are both transcription factors involved in the
235
regulation of hematopoiesis and vasculogenesis (35). A statistically significant decrease
236
in the level of fli1a gene transcripts was observed from 12 hpi. Similar results were
237
observed for ets1a, as the time course profile for this transcription factor was identical
238
to that for fli1a. The interindividual variations in ve-cad, a key regulator of endothelial
239
intercellular junctions (36), made it difficult to observe a clear expression pattern,
240
particularly at the late stages of the infection. This effect might be associated with the
241
observed heterogeneity in the appearance of hemorrhages, in terms of time, area or
242
larvae. Notably, the downregulation of ve-cad at 9 hpi suggests an increase in the
243
permeability of the endothelium at that infection point. Interestingly, fli1a
244
downregulation clearly occurred before the onset of hemorrhages and the loss of
245
circulation, suggesting that hemorrhages were a consequence of this downregulation.
246
Zebrafish antiviral response to SVCV. To further study the mechanisms involved in
247
the defense against SVCV, we analyzed the expression of some important genes
248
associated with viral detection. Pattern recognition receptors, such as Toll-like receptors
249
(TLRs) and RIG-1-like receptors (RLRs), are membrane-bound and cytosolic sensors,
250
respectively (37). We observed a significant downregulation of tlr7, tlr22, tlr8a and
251
mda5 at 3 hpi (Fig. 3A). As occurred with Dengue virus (38) or Influenza (39), this
252
downmodulation could correspond to an immune evasion strategy. rig-1 was not
253
regulated during the first 24 hours of the infection. Regarding tlr3, we also saw a slight,
254
but significant, regulation at 12 hpi.
12
255
Pattern recognition receptors initiate antimicrobial defense mechanisms through several
256
conserved signaling pathways (40). The activation of interferon-regulatory factors
257
induces the production of several antiviral response-related genes. We observed that
258
SVCV elicited a response in all the analyzed genes (Fig. 3B). ifnphi1 and ifnphi2
259
showed different response profiles. ifnphi1 remained downregulated between 3 and 9
260
hpi, the time in which ifnphi2 was regulated. ifnphi1 was upregulated from 12 hpi,
261
similar to mx (a and b paralogues), which increased from 12 hours and reached a
262
maximum at 24 hours. rarres3 showed a biphasic response, with higher expression at 9
263
and 24 hpi. The recently characterized ifit gene family in zebrafish (41, 42) was also
264
present in the response, highlighting the response of ifit17a, which reached a 20-fold
265
change.
266
SVCV positive cells are not endothelial cells. To determine whether endothelial cells
267
were the primary targets of SVCV in this in vivo model of infection we performed an
268
immunohistochemistry using an antibody against SVCV (Fig. 4). Using Fli1a-GFP
269
infected larvae, SVCV was clearly visible in areas surrounding vessels (Fig. 4A). The
270
3D reconstruction and volume clipping of the Fli1a channel showed that the SVCV
271
positive cells were not endothelial cells (Fig. 4B, C, D). We observed crossing and
272
circulating SVCV-positive cells, suggesting that these cells might be leukocytes
273
interacting with the endothelium.
274
Our hypothesis was confirmed by performing L-plastin labelling (Fig. 5). L-
275
plastin positive cells were observed surrounding vessels and many of these cells co-
276
localized with SVCV. Again, the 3D reconstruction of the confocal images conferred
277
information about the cells and their position relative to the vessels (Fig. 5A). In
278
addition, we observed that SVCV-infected cells showed viral antigen positive signal
13
279
inside and on their surface (Fig. 5A). The signal detected on the cell surface might be
280
associated with viral assembly, antigen presentation by infected cells or both (6). We
281
observed cells labeled with only L-plastin (interacting or not with the endothelium),
282
double-positive cells (interacting or not with the endothelium) and cellular debris
283
positive for one or both markers. Large amounts of cellular debris were observed,
284
always with L-plastin-positive cells phagocytising SVCV-positive particles in the same
285
area (Fig. 5B). This cellular debris might play a role in the first line of defense against
286
SVCV, as observed with other pathogens (43). Furthermore, defective debris clearance
287
has been associated with persistent inflammatory diseases (44).
288
Cellular response to SVCV infection. The transendothelial migration of leukocytes
289
during inflammation has been well characterized (45). In the present study, the role of
290
leukocytes during infection was explored using the Mpx-GFP zebrafish transgenic line
291
combined with L-plastin immunohistochemistry (Fig. 6). Using this system, we were
292
able to differentiate between macrophages (L-plastin+, Mpx-) and neutrophils (L-
293
plastin+, Mpx+). After 24 hours of infection, practically all macrophages were
294
undetectable (Fig. 6A) and a statistically significant migration of neutrophils to the head
295
was observed (Fig. 6A, 6B, 6C). The loss of leukocytes in the caudal hematopoietic
296
region was also confirmed (Fig. 6B). Using markers for both cellular types, marco for
297
macrophages and mpx for neutrophils, we could check by qPCR the dynamics of that
298
leukocytes population (Fig. 6D). The analysis of marco gene expression suggested an
299
initial decrease in the macrophage population between 3 and 6 hpi and then,
300
macrophages remained stable until 18 hpi. Thereafter, the macrophage population
301
markedly reduced to nearly 0 at 24 hpi (Fig. 6D). Regarding neutrophils, mpx revealed
302
no differences in the number of cells, thus confirming the migration observed in
14
303
confocal images of these cells during the infection. Complementary experiments using
304
other transgenic lines possessing fluorescent macrophages (Mpeg-GFP), neutrophils
305
(Mpx-GFP) or thrombocytes (Cd41-GFP) showed that only macrophages were SVCV
306
positive at 18 hpi (data not shown). Moreover, given the importance of platelets and
307
coagulation system in hemorrhagic infections (46) we used the transgenic line with
308
fluorescent thrombocytes (fish equivalent to mammalian platelets) to detect a possible
309
thrombocytopenia or coagulopathy after SVCV infection. As no differences were found
310
in thrombocyte behavior or count between control and SVCV-infected larvae (data not
311
shown), we decided to focus our attention in macrophages.
312
To further confirm macrophage role as preferred target for this hemorrhagic
313
virus, at least during the first hours of the infection, we took advantage of the
314
morpholino-mediated gene knockdown technique. We injected the morpholino (MO)
315
PU.1 at a concentration of 1 mM to prevent the formation of macrophages in embryos
316
(47). A comparison of the normal fish and macrophage-depleted fish in these
317
experimental infections showed a delay in the mortality of morphants (Fig. 6E),
318
although eventually the mortality reached 100% in both groups. The duration of the MO
319
effect was analyzed at an equivalent time of 48 hpi in non-infected fish. Marco
320
expression was completely inhibited, as demonstrated using qPCR analysis (Fig. 6E). In
321
contrast, the population of neutrophils was not affected by the morpholino, as
322
demonstrated by the mpx gene expression after 48 hpi (Fig. 6E). The approximate 9-
323
hour delay in the mortality between both groups, suggested that macrophage presence is
324
crucial for SVCV pathogenesis during the first hours of infection. Therefore, we
325
analyzed the transcription levels of the SVCV N gene at 9 hpi (Fig. 6E). At this time,
326
fish without macrophages had lower viral transcription levels compared with normal
15
327
fish. Macrophage depletion delays the kinetics of viral transcription but is not enough to
328
prevent the death of morphant fish. This effect likely reflects the high virulence of the
329
infection. As an obligate intracellular pathogen, SVCV infects to survive and even if the
330
primary target is not present, the virus could still infect other cells. However, the
331
efficiency of infection is reduced in the first hours, resulting in a different timing of
332
mortality. This, plus the fact that the only SVCV-positive cells detected at 18 hpi are
333
macrophages, suggested that these cells could be the primary target of the virus.
334
SVCV elicits an inflammatory response that induces macrophages pyroptosis and
335
IL1β release. The fact that macrophages were primarily infected with SVCV prompted
336
us to examine the “disappearance” of these cells and determine whether pyroptosis is
337
involved. Pyroptosis might play a role in pathogen clearance, leading to inflammatory
338
pathology, which is detrimental to the host but positively affects the spread of the
339
pathogen (3). This programmed cell death depends on inflammasome activation and
340
consequent Caspase-1 action (48). In zebrafish, Caspa associates with ASC, an
341
inflammasome adaptor (49). Moreover, it has recently been demonstrated that this
342
caspase is able to cleave Il1 in zebrafish (50). Using qPCR, we characterized the
343
regulation of some important genes involved in the inflammatory response (Fig. 7A).
344
Three typical pro-inflammatory cytokines, il1, tnf and il6, were regulated during
345
infection. Il1 was the first responder, and a clear biphasic profile was observed. The
346
initial up-regulation of cytokine expression might be associated with the recognition of
347
SVCV by TLRs, followed by a second up-regulation of cytokine with a consequent
348
inflammatory response (51). This inflammatory response also triggered the response of
349
Tnf and Il6. Tnf was induced from 12 to 24 hpi. In addition, il6 showed a subtle
350
expression peak at 18 hpi. Strikingly, the response of this cytokine was delayed and
16
351
weak. Similar results were observed in the caspa time course. A complete inhibition of
352
asc was observed at 9 hpi, then, the expression of asc reached a maximum at 18 hpi
353
(Fig. 7A).
354
Immunohistochemistry against Caspa and SVCV showed that the macrophages
355
remaining at 20 hpi were positive for Caspa (Fig. 7B). These macrophages showed
356
interactions with infected cells (Fig. 7B). In addition, Mpeg1-positive and Caspa-
357
positive vesicles, presumably from macrophages, were observed in the same area near
358
SVCV-positive cells. Consistently with previus experiments, the 3D reconstruction
359
revealed that infected macrophages had external SVCV-positive signals (Fig. 7B).
360
Significant differences were found in Caspa levels between control and infected fish
361
(Fig 7C) showing the participation of that protein in the infection.
362
Il1 regulation is essential for a proper acute inflammatory response to an
363
infectious challenge. The implication of Caspa in Il1 secretion prompted us to
364
investigate the presence of that pro-inflammatory cytokine in infected larvae. An
365
antibody against zebrafish Il1 in combination with the Mpeg1-GFP transgenic line
366
allowed us the detection of this cytokine in macrophages (Fig. 7D). A 3D reconstruction
367
of the confocal images was generated to amplify the details, and two types of Il1-
368
positive macrophages were observed in SVCV-infected larvae. The first type was
369
SVCV-positive macrophages (Fig. 7E). The second type of macrophages localized near
370
zones with SVCV signal, was Il1 positive but was not positive for the virus (Fig. 7F).
371
The difference between these two macrophages types was the amount of Il1 detected.
372
Non-permissive macrophages in areas with SVCV-positive signal contained a higher
373
amount of Il1. Furthermore, the Mpeg1 signal was lost in favor of Il1 signal, which
374
could lead to the release of this cytokine. Il1 was detected in the macrophages of non17
375
infected fish, but the signal in these resting macrophages was clearly different that
376
observed in stimulated cells (Fig. 8A). Macrophages from infected fish showed a higher
377
amount of Il1 and what could be the release of this cytokine by microvesicle shedding
378
(Fig. 8B) or directly through the cell membrane (Fig. 8C) were observed. To assess the
379
membrane disruption produced during pyroptosis we used a cell membrane-impermeant
380
dye, PI. This dye was used in vivo and the uptake of PI by macrophages during SVCV
381
infection was confirmed (Fig. 8D). Furthermore, TUNEL assay showed IL1β releasing
382
cells dying during infection (Fig 8E).
383
18
384
DISCUSSION
385
In the present work we obtained images of the infection at the single cell level
386
using a zebrafish model, identifying macrophages as the preferred target of SVCV. We
387
were able to demonstrate for the first time that pyroptosis and Il1β release were
388
involved in the viral-induced cell dead in a whole organism. The direct observation of
389
such processes in individual cells provided a deeper knowledge of the inflammatory
390
mechanisms implicated in this viral disease.
391
Hemorrhagic viral diseases are globally observed with important pathogens,
392
such as Dengue Virus or Hantaviruses (52). Although hemorrhages have been
393
associated with the infection of endothelial cells (1), other cellular types could be the
394
main targets for viral infection (4). The lack of adequate in vivo infection models has
395
limited the research on viral pathogenesis. In SVCV-infected zebrafish larvae we
396
observed a loss of fluorescence of the endothelial marker Fli1a. However, this was not
397
due to infection of endothelial cells or cytopathic effects, and vascular integrity
398
remained intact, at least in the tail. The fact that haemorrhages always occur in the head
399
of the larva suggests the possibility that these occur as a result of damage or
400
permeability regulation in a particular kind of vessel, as it is known that the control of
401
the permeability may be different in different vessel types (53). The dysregulation of the
402
vascular endothelium, observed as a loss of Fli1a fluorescence, could reflect strong
403
immune activation during infection. In addition, the response induced in the rest of the
404
organism should also be considered. The downregulation of important transcription
405
factors, such as Fli1a and Ets1a, in endothelial cells shows the importance of a response
406
of the organism as a whole. It has been shown that the downregulation of Fli1 in
407
response to LPS is mediated through TLRs (54). Endothelial cells express different
19
408
receptors, such as Tlr3, Rig-1 and Mda5, among others. Based on this fact and the
409
qPCR results obtained in the present study, the activation of endothelial cells through
410
the Tlr3 might the cause of the loss of Fli1a expression, thereby inducing loss of
411
circulation, hemorrhages and inhibition in the development of hematopoietic precursors.
412
Recent studies have shown a relationship between Fli1 and glycosphingolipids
413
metabolism (55). Thus, we could consider Fli1a downregulation as a mechanism for
414
protection against viral infection in endothelial cells. Although, similar to other
415
hemorrhagic viral infections, we cannot rule out the possibility that endothelial cells
416
could be infected during the final stages of infection (4). However, this situation would
417
not be decisive for the pathogenesis of the disease.
418
Altogether, the qPCR results provide interesting data regarding the dynamics of
419
the response. The viral transcription began to emerge from 9 hpi, time in which we
420
observed the first increase in il1 transcription and the strong downregulation of asc.
421
With regard to the complete inhibition of asc transcription observed at 9 hpi, we cannot
422
rule out the involvement of the virus. Pathogens have developed diverse strategies to
423
block some of the components of the inflammasome (56). It has been shown that ASC
424
transcription is downregulated during the infection of keratinocytes with some types of
425
human papillomavirus (57). The host might also control this response (58); thus, it is
426
likely that, in this case, the host could control macrophage loss through the inhibition of
427
asc transcription. However, the potential blockade by the virus to favor viral spread
428
cannot be excluded. Furthermore, the accumulation of Caspa as an unprocessed protein
429
(pro-caspase) into the cells (59) would explain the weak transcriptional regulation
430
observed for this gene during the infection.
20
431
Recent studies have demonstrated the importance of the inflammasome in Il1
432
generation upon infection with RNA viruses (60) and showed that a biphasic response
433
might be associated with the two signals necessary for Il1 release from macrophages
434
(61). If we take into account of the results obtained also at protein level, the second
435
upregulation observed at 12 hpi might be related with inflammasome activation.
436
Inflammasome activation could be induced directly by the virus or cellular debris, but
437
further investigation is necessary to confirm the type or types of inflammasomes
438
implicated in this process. Moreover, as occurs with Influenza virus (62), and based in
439
the obtained qPCR results, we can not exclude the involvement of Tlr7 in
440
inflammasome activation and IL1β release.
441
The zebrafish model facilitated the capture of images showing Il1 release from
442
single macrophages during SVCV infection. The shedding of microvesicles containing
443
fully processed Il1 from the plasma membrane represents a major secretory pathway
444
for the rapid release of this cytokine (63). We observed vesicles presumably derived
445
from macrophages in areas with cellular debris or SVCV-positive cells. However, we
446
also observed the possible direct release of Il1 through the cellular membrane. Rarres3,
447
more commonly known as an antiviral protein, has recently been associated with
448
increased microtubule stability in cells (64), and cells with more stable microtubules
449
displayed increased pyroptosis (65). The fact that the expression profile of rarres3
450
overlaps with that of Il1 should be addressed in future studies.
451
The transcriptional regulation of some of the genes associated with pyroptosis
452
and the complete disappearance of macrophages before the onset of mortality, being
453
positive for Caspa and Il1β, suggested that pyroptosis is involved in SVCV infection. In
454
recent years, there have been many advances in research concerning this type of cell
21
455
death, particularly with regard to bacteria (66). In the case of viruses, our understanding
456
remains a step behind. It is known that some viruses cause pyroptosis of macrophages in
457
vitro, but so far, it is less clear whether this process occurs in vivo (67, 68). Using the
458
zebrafish model, we have provided the first description of viral-induced pyroptosis in
459
the context of a whole organism. The fact that a hemorrhagic viral infection was able to
460
induce macrophage pyroptosis in zebrafish, in the context of whole larvae, promotes the
461
use of this model organism for studying host-pathogen interactions.
462
Future studies will be particularly important for an in deep study of the role of
463
the endothelium, the inflammasome and pyroptosis in defense against the infection.
464
Indeed, without losing sight of the possible effects that the virus might have on the
465
regulation of the host response, the results obtained in the present study could be the
466
starting point to understand the antiviral immune response as a whole. Furthermore, the
467
application of these techniques and studies to other pathogens will improve the current
468
knowledge of host-pathogen interactions and increase the potential for the discovery of
469
new therapeutic targets.
470
22
471
ACKNOWLEDGMENTS
472
This work was supported by the projects CSD2007-00002 “Aquagenomics” and
473
AGL2011-28921-C03 from the Spanish Ministerio de Economía e Innovación and
474
European structural funds (FEDER) / Ministerio de Ciencia e Innovación (CSIC08-1E-
475
102). Our laboratory is also funded by ITN 289209 “FISHFORPHARMA” (EU) and
476
project 201230E057 from the Agencia Estatal Consejo Superior de Investigaciones
477
Científicas (CSIC).
478
The authors would like to thank Rubén Chamorro for technical assistance,
479
Professor Paul Martin (University of Bristol) for the provision of the L-plastin antibody
480
and Dr. John Hansen (Western Fisheries Research Center) for the provision of the Il1
481
antibody. Mónica Varela received predoctoral grants from the JAE Program (CSIC and
482
European structural Funds).
483
23
484
REFERENCES
485
1. Valbuena G, Walker DH. 2006. The endothelium as a target for infections. Annu.
486
Rev. Pathol. 1:171-198.
487
2. Muller WA. 2003. Leukocyte-endothelial cell interactions in leukocyte
488
transmigration and the inflammatory response. Trends. Immunol. 24:327-334.
489
3. Lupfer CR, Kanneganti TD. 2012. The role of inflammasome modulation in
490
virulence. Virulence 3:262-270.
491
4. Paessler S, Walker DH. 2013. Pathogenesis of the viral Hemorrhagic Fevers. Annu.
492
Rev. Pathol. 8:411-440.
493
5. Bray M, Geisbert TW. 2005. Ebola virus: the role of macrophages and dendritic
494
cells in the pathogenesis of Ebola hemorrhagic fever. Int. J. Biochem. Cell. Biol.
495
37:1560–1566.
496
6. Halstead SB, O´Rourke EJ, Allison AC. 1977. Dengue viruses and mononuclear
497
phagocytes. II. Identity of blood and tissue leukocytes supporting in vitro infection. J.
498
Exp. Med. 146: 218-228.
499
7. Aye KS, Charngkaew K, Win N, Wai KZ, Moe K, Punyadee N, Thiemmeca S,
500
Suttitheptumrong A, Sukpanichnant S, Prida M, Halstead SB. 2014. Pathologic
501
highlights of dengue hemorrhagic fever in 13 autopsy cases from Myanmar. Hum.
502
Pathol. 45:1221-33
503
8. Mercer J, Greber UF. 2013. Virus interactions with endocytic pathways in
504
macrophages and dendritic cells. Trends Microbiol. 21:380-388.
24
505
9. Howe K, Clark MD, Torroja CF, Torrance J, Berthelot C, Muffato M, Collins
506
JE, Humphray S, McLaren K, Matthews L, McLaren S, Sealy I, Caccamo M,
507
Churcher C, Scott C, Barrett JC, Koch R, Rauch GJ, White S, Chow W, Kilian B,
508
Quintais LT, Guerra-Assunção JA, Zhou Y, Gu Y, Yen J, Vogel JH, Eyre T,
509
Redmond S, Banerjee R, Chi J, Fu B, Langley E, Maguire SF, Laird GK, Lloyd D,
510
Kenyon E, Donaldson S, Sehra H, Almeida-King J, Loveland J, Trevanion S, Jones
511
M, Quail M, Willey D, Hunt A, Burton J, Sims S, McLay K, Plumb B, Davis J,
512
Clee C, Oliver K, Clark R, Riddle C, Elliot D, Threadgold G, Harden G, Ware D,
513
Begum S, Mortimore B, Kerry G, Heath P, Phillimore B, Tracey A, Corby N,
514
Dunn M, Johnson C, Wood J, Clark S, Pelan S, Griffiths G, Smith M, Glithero R,
515
Howden P, Barker N, Lloyd C, Stevens C, Harley J, Holt K, Panagiotidis G, Lovell
516
J, Beasley H, Henderson C, Gordon D, Auger K, Wright D, Collins J, Raisen C,
517
Dyer L, Leung K, Robertson L, Ambridge K, Leongamornlert D, McGuire S,
518
Gilderthorp R, Griffiths C, Manthravadi D, Nichol S, Barker G, Whitehead S, Kay
519
M, Brown J, Murnane C, Gray E, Humphries M, Sycamore N, Barker D,
520
Saunders D, Wallis J, Babbage A, Hammond S, Mashreghi-Mohammadi M, Barr
521
L, Martin S, Wray P, Ellington A, Matthews N, Ellwood M, Woodmansey R, Clark
522
G, Cooper J, Tromans A, Grafham D, Skuce C, Pandian R, Andrews R, Harrison
523
E, Kimberley A, Garnett J, Fosker N, Hall R, Garner P, Kelly D, Bird C, Palmer S,
524
Gehring I, Berger A, Dooley CM, Ersan-Ürün Z, Eser C, Geiger H, Geisler M,
525
Karotki L, Kirn A, Konantz J, Konantz M, Oberländer M, Rudolph-Geiger S,
526
Teucke M, Lanz C, Raddatz G, Osoegawa K, Zhu B, Rapp A, Widaa S, Langford
527
C, Yang F, Schuster SC, Carter NP, Harrow J, Ning Z, Herrero J, Searle SM,
528
Enright A, Geisler R, Plasterk RH, Lee C, Westerfield M, de Jong PJ, Zon LI,
529
Postlethwait JH, Nüsslein-Volhard C, Hubbard TJ, Roest Crollius H, Rogers J,
25
530
Stemple DL. 2013. The zebrafish reference genome sequence and its relationship to the
531
human genome. Nature 496:498-503.
532
10. Crim MJ, Riley LK. 2012. Viral diseases in zebrafish: what is known and
533
unknown. ILAR J. 53:135-143.
534
11. Encinas P, Rodriguez-Milla MA, Novoa B, Estepa A, Figueras A, Coll, J. 2010.
535
Zebrafish fin immune responses during high mortality infections with viral
536
haemorrhagic septicemia rhabdovirus. A proteomic and transcriptomic approach. BMC
537
Genomics 11:518.
538
12. López-Muñoz A, Roca FJ, Sepulcre MP, Meseguer J, Mulero V. 2010. Zebrafish
539
larvae are unable to mount a protective antiviral response against waterborne infection
540
by spring viremia of carp virus. Dev. Comp. Immunol. 34:546–552.
541
13. Ludwig M, Palha N, Torhy C, Briolat V, Colucci-Guyon E, Brémont M,
542
Herbomel P, Boudinot P, Levraud, JP. 2011. Whole-body analysis of a viral
543
infection: vascular endothelium is a primary target of infectious hematopoietic necrosis
544
virus in zebrafish. PLoS Pathog. 7:e1001269.
545
14. Burgos JS, Ripoll-Gomez J, Alfaro JM, Sastre I, Valdivieso F. 2008. Zebrafish
546
as a new model for herpes simplex virus type 1 infection. Zebrafish 5:323-333.
547
15. Hubbard S, Darmani NA, Thrush GR, Dey D, Burnham L, Thompson JM,
548
Jones K, Tiwari V. 2010. Zebrafish-encoded 3-O-sulfotransferase-3 isoform mediates
549
herpes simplex virus type 1 entry and spread. Zebrafish 7:181-187.
550
16. Palha N, Guivel-Benhassine F, Briolat V, Lutfalla G, Sourisseau M, Ellett F,
551
Wang CH, Lieschke GJ, Herbomel P, Schwartz O, Levraud JP. 2013. Real-time
26
552
whole-body visualization of Chikungunya Virus infection and host interferon response
553
in zebrafish. PLoS Pathog. 9:e1003619.
554
17. Ahne W, Bjorklund HV, Essbauer S, Fijan N, Kurath G, Winton JR. 2002.
555
Spring viraemia of carp (SVC). Dis. Aquat. Organ. 52:261–272.
556
18. Taylor NG, Peeler EJ, Denham KL, Crane CN, Thrush MA, Dixon PF, Stone
557
DM, Way K, Oidtman, BC. 2013. Spring viraemia of carp (SVC) in the UK: The road
558
to freedom. Prev. Vet. Med. 111:156-164.
559
19. Westerfield M. 2000. The zebrafish book. A guide for the laboratory use of
560
zebrafish (Danio rerio). Eugene: University of Oregon Press.
561
20. Nüsslein-Volhard C, Dahm R. 2002. Zebrafish: a practical approach. New York:
562
Oxford University Press. 303 pp.
563
21. Reed JL, Muench H. 1938. A simple method of estimating fifty per cent end point.
564
Am. J. Hyg. 27:493-497.
565
22. Benard EL, van der Sar AM, Ellett F, Lieschke GJ, Spaink HP, Meijer AH.
566
2012. Infection of zebrafish embryos with intracellular bacterial pathogens. J. Vis. Exp.
567
61:e3781.
568
23. Cui C, Benard EL, Kanwal Z, Stockhammer OW, van der Vaart M,
569
Zakrzewska A, Spaink HP, Meijer AH. 2011. Infectious disease modeling and innate
570
immune function in zebrafish embryos. Methods Cell Biol. 105:273-308.
571
24. Varela M, Dios S, Novoa B, Figueras A. 2012. Characterisation, expression and
572
ontogeny of interleukin-6 and its receptors in zebrafish (Danio rerio). Dev. Comp.
573
Immunol. 37:97-106.
27
574
25. Espín R, Roca FJ, Candel S, Sepulcre, MP, González-Rosa, JM. 2013. TNF
575
receptors regulate vascular homeostasis in zebrafish through a caspase-8, caspase-2 and
576
P53 apoptotic program that bypasses caspase-3. Dis. Model. Mech. 6:383-396.
577
26. Rhodes J, Hagen A, Hsu K, Deng M, Liu TX, Look AT, Kanki JP. 2005.
578
Interplay of pu.1 and gata1 determines myelo-erythroid progenitor cell fate in zebrafish.
579
Dev. Cell 8:97–108.
580
27. Levraud JP, Boudinot P, Colin I, Benmansour A, Peyrieras N, Herbomel P,
581
Lutfalla G. 2007. Identification of the zebrafish IFN receptor: implications for the
582
origin of the vertebrate IFN system. J. Immunol. 178:4385-4394.
583
28. Lawson ND, Weinstein BM. 2002. In vivo imaging of embryonic vascular
584
development using transgenic zebrafish. Dev. Biol. 248:307-318.
585
29. Mathias JR, Dodd ME, Walters KB, Yoo SK, Ranheim EA, Huttenlocher A.
586
2009. Characterization of zebrafish larval inflammatory macrophages. Dev. Comp.
587
Immunol. 33:1212-1217.
588
30. Herbomel P, Thisse B, Thisse C. 1999. Ontogeny and behaviour of early
589
macrophages in the zebrafish embryo. Development 126:3735-3745.
590
31. Brannon MK, Davis JM, Mathias JR, Hall CJ, Emerson JC, Crosier PS,
591
Huttenlocher A, Ramakrishnan L, and Moskowitz SM. 2009. Pseudomonas
592
aeruginosa Type III secretion system interacts with phagocytes to modulate systemic
593
infection of zebrafish embryos. Cell Microbiol. 11:755-768.
28
594
32. Wang L, Zhang P, We, Y, Gao Y, Patient R, Liu F. 2011. A blood flow–
595
dependent klf2a-NO signaling cascade is required for stabilization of hematopoietic
596
stem cell programming in zebrafish embryos. Blood 118:4102-4110.
597
33. Spyropoulos DD, Pharr PN, Lavenburg KR, Jackers P, Papas TS, Ogawa M,
598
Watson DK. 2000. Hemorrhage, impaired hematopoiesis, and lethality in mouse
599
embryos carrying a targeted disruption of the Fli1 transcription factor. Mol. Cell. Biol.
600
20:5643–5652.
601
34. Pham VN, Lawson ND, Mugford JW, Dye L, Castranova D, Lo B, Weinstein,
602
BM. 2007. Combinatorial function of ETS transcription factors in the developing
603
vasculature. Dev. Biol. 303:772–783.
604
35. Sumanas S, Lin S. 2006. Ets1-Related Protein Is a Key Regulator of
605
Vasculogenesis in Zebrafish. PLoS Biol. 4:e10.
606
36. Vestweber D. 2008. VE-cadherin: the major endothelial adhesion molecule
607
controlling cellular junctions and blood vessel formation. Arterioscler. Thromb. Vasc.
608
Biol. 28:223-232.
609
37. Kawai T, Akira S. 2006. Innate immune recognition of viral infection. Nat.
610
immunol. 7:131-137.
611
38. Chang TH, Chen SR, Yu CY, Lin YS, Chen YS, Kubota T, Matsuoka M, Lin
612
YL. 2012. Dengue Virus Serotype 2 Blocks Extracellular Signal-Regulated Kinase and
613
Nuclear Factor-κB Activation to Downregulate Cytokine Production. PLoS One
614
7:e41635.
29
615
39. Keynan Y, Fowke KR, Ball TB, Meyers, AFA. 2011. Toll-Like Receptors
616
Dysregulation after Influenza Virus Infection: Insights into Pathogenesis of Subsequent
617
Bacterial Pneumonia. ISRN Pulmonology 2011:142518.
618
40. Broz P, Monack DM. 2013. Newly described pattern recognition receptors team up
619
against intracellular pathogens. Nat. Immunol. Rev. 13:551-565.
620
41. Liu Y, Zhang YB, Liu TK, Gui JF. 2013. Lineage-Specific Expansion of IFIT
621
Gene Family: An Insight into Coevolution with IFN Gene Family. PLoS One 8:e66859.
622
42. Varela M, Díaz-Rosales P, Pereiro P, Forn-Cuní G, Costa MM, Dios S, Romero
623
A, Figueras A, Novoa B. 2014. Interferon-induced genes of the expanded IFIT family
624
show conserved antiviral activities in non-mammalian species. PLoS One (in press).
625
43. Norling LV, Perretti M. 2013. Control of Myeloid Cell Trafficking in Resolution.
626
J. Innate. Immun. 5:367-376.
627
44. Nathan C, Ding A. 2010. Nonresolving Inflammation. Cell 140:871–882.
628
45. Muller WA. 2009. Mechanisms of Transendothelial Migration of Leukocytes. Circ.
629
Res. 105:223-230.
630
46. Zapata JC, Cox D, Salvato MS. 2014. The Role of Platelets in the Pathogenesis of
631
Viral Hemorrhagic Fevers. PLoS Negl. Trop. Dis. 8:e2858.
632
47. He S, Lamers GE, Beenakker JW, Cui C, Ghotra VP, Danen EH, Meijer AH,
633
Spaink HP, Snaar-Jagalska BE. 2012. Neutrophil-mediated experimental metastasis is
634
enhanced by VEGFR inhibition in a zebrafish xenograft model. J. Pathol. 227:431–445.
635
48. Miao EA, Rajan JV, Aderem A. 2011. Caspase-1-induced Pyroptotic cell death.
636
Immunol. Rev. 243:206-214.
30
637
49. Masumoto J, Zhou W, Chen FF, Su F, Kuwada JY. 2003. Caspy, a Zebrafish
638
Caspase, Activated by ASC Oligomerization Is Required for Pharyngeal Arch
639
Development. J. Biol. Chem. 278:4268-4276.
640
50. Vojtech LN, Scharping N, Woodson JC, Hansen JD. 2012. Roles of
641
inflammatory caspases during processing of zebrafish interleukin-1β in Francisella
642
noatunensis infection. Infect. Immun. 80:2878-85.
643
51. Negash AA, Ramos HJ, Crochet N, Lau DT, Doehle B, Papic N, Delker DA, Jo
644
J, Bertoletti A, Hagedorn CH, Gale MJ. 2013. IL-1β Production through the NLRP3
645
Inflammasome by Hepatic Macrophages Links Hepatitis C Virus Infection with Liver
646
Inflammation and Disease. PLoS Pathog. 9:e1003330.
647
52. Cox D, Salvato MS, Zapata JC. 2013. The role of platelets in viral hemorrhagic
648
fevers. J. Bioterr. Biodef. S12:2.
649
53. Dejana E, Orsenigo F. 2013. Endothelial adherens junctions at a glance. J. Cell Sci.
650
126:2545-2549.
651
54. Ho HH, Ivashkiv LB. 2010. Downregulation of friend leukemia virus integration 1
652
as a feedback mechanism that restrains lipopolysaccharide induction of matrix
653
metalloproteases and interleukin-10 in human macrophages. J. Interferon Cytokine Res.
654
30:893-900.
655
55. Richard EM, Thiyagarajan T, Bunni MA, Basher F, Roddy PO, Siskind LJ,
656
Nietert PJ, Nowling TK. 2013. Reducing FLI1 Levels in the MRL/lpr Lupus Mouse
657
Model Impacts T Cell Function by Modulating Glycosphingolipid Metabolism. PLoS
658
One 8:e75175.
31
659
56. Taxman DJ, Huang MT, Ting JP. 2011. Inflammasome Inhibition as a Pathogenic
660
Stealth Mechanism. Cell Host Microbe 8:7-11.
661
57. Karim R, Meyers C, Backendorf C, Ludigs K, Offringa R, van Ommen GJ,
662
Melief CJ, van der Burg SH, Boer, JM. 2011. Human Papillomavirus Deregulates the
663
Response of a Cellular Network Comprising of Chemotactic and Proinflammatory
664
Genes. PLoS One 6:e17848.
665
58. Bedoya F, Sandler LL, Harton JA. 2007. Pyrin-only protein 2 modulates NF-
666
kappaB and disrupts ASC:CLR interactions. J. Immunol. 178:3837–3845.
667
59. Fernandes-Alnemri T, Wu J, Yu JW, Datta P, Miller B, Jankowski W,
668
Rosenberg S, Zhang J, Alnemri ES. 2007. The pyroptosome: a supramolecular
669
assembly of ASC dimers mediating inflammatory cell death via caspase-1 activation.
670
Cell Death Differ. 14:1590-1604.
671
60. Poeck H, Bscheider M, Gross O, Finger K, Roth S, Rebsamen M,
672
Hannesschläger N, Schlee M, Rothenfusser S, Barchet W, Kato H, Akira S, Inoue
673
S, Endres S, Peschel C, Hartmann G, Hornung V, Ruland J. 2010. Recognition of
674
RNA virus by RIG-I results in activation of CARD9 and inflammasome signaling for
675
interleukin 1 beta production. Nat. Immunol. 11:63-69.
676
61. Netea MG, Simon A, van de Veerdonk F, Kullberg BJ, van der Meer JW,
677
Joosten LA. 2010. IL-1β processing in host defense: beyond the inflammasomes. PLoS
678
Pathog. 6:e1000661.
679
62. Ichinohe T, Pang IK, Iwasaki A. 2010. Influenza virus activates inflammasomes
680
via its intracellular M2 ion channel. Nat. Immunol. 11:404-410.
32
681
63. Piccioli P, Rubartelli A. 2013. The secretion of IL-1 and options for release.
682
Semin. Immunol. 25:425-9.
683
64. Scharadin TM, Jiang H, Martin S, Eckert RL. 2012. TIG interaction at the
684
centrosome alters microtubule distribution and centrosome function. J. Cell. Sci.
685
125:2604-2614.
686
65. Salinas RE, Ogohara C, Thomas MI, Shukla KP, Miller SI, Ko DC. 2014. A
687
cellular genome-wide association study reveals human variation in microtubule stability
688
and a role in inflammatory cell death. Mol. Biol. Cell 25:76-86.
689
66. Richard EM, Thiyagarajan T, Bunni MA, Basher F, Roddy PO, Sarkar A,
690
Warren SE, Wewers MD, Aderem A. 2010. Caspase-1-induced pyroptosis is an
691
innate immune effector mechanism against intracellular bacteria. Nat. Immunol.
692
11:1136-1142.
693
67. Aachoui Y, Sagulenko V, Miao EA, Stacey KJ. 2013. Inflammasome-mediated
694
pyroptotic and apoptotic cell death, and defense against infection. Curr. Opin.
695
Microbiol. 16:319-326.
696
68. Tan TY, Chu JJ. 2013. Dengue virus-infected human monocytes trigger late
697
activation of caspase-1, which mediates pro-inflammatory IL-1β secretion and
698
pyroptosis. J. Gen. Virol. 94:1215-1220.
699
33
700
FIGURE LEGENDS
701
Figure 1. Infection course and loss of endothelium fluorescence in infected larvae
702
(A) Kaplan-Meier survival curve after infection with 2 nl of a SVCV solution of 3x106
703
TCID50/ml (p<0.001). (B) Quantification of SVCV N gene transcripts as a measure of
704
the viral amount in the larvae over time. Heat-killed SVCV was injected in order to
705
separate immune response against the initial viral load. At 3 hpi significant differences
706
in the quantification of SVCV N gene transcripts were found between heat-killed SVCV
707
and SVCV infected larvae. The samples were obtained from infected larvae at different
708
times post-infection. Each sample (technical triplicates) was normalized to the 18S
709
gene. To avoid differences due to the larval development stage, the samples were
710
standardized with respect to an uninfected control of the same age. Three biological
711
replicates (of 10 larvae each) were used to calculate the means ± standard error of mean.
712
Significant differences between infected and non-infected larvae at the same time point
713
were displayed as *** (0.0001<p<0.001), ** (0.001<p<0.01) or * (0.01<p<0.05). qPCR
714
results are representative of 2 independent experiments. (C) The common macroscopic
715
symptoms observed at 24 hpi were hemorrhages and blood accumulation around the
716
eyes and head (arrows). See also Movie S1-S3 in supplemental material. (D). Zebrafish
717
larvae head confocal images (maximal projections from multiple z-stacks) of infected 3
718
dpf Tg(Fli1a-GFP) larvae immunostained with L-plastin antibody at 24 hpi (20x
719
magnification) showing the loss of GFP fluorescence in infected fish. (E) Confocal
720
images (maximal projections from multiple z-stacks) of infected 3 dpf Tg(Fli1a-GFP)
721
larvae immunostained with L-plastin antibody at 24 hpi (20x magnification) of vessels
722
in the yolk sac and trunk showing the loss of GFP fluorescence in infected fish. (F)
34
723
Higher magnification images are shown morphological changes in L-plastin-positive
724
cells in SVCV-infected fish.
725
Figure 2. Endothelium integrity is conserved 24 hours post-infection. (A)
726
Microangiography performed at 24 hpi injecting tetramethylrhodamine into the caudal
727
vein of 3 dpf infected and control larvae. Tetramethylrhodamine remained inside
728
vessels, even in infected fish without Fli1a-GFP fluorescence. Higher magnification
729
images are shown in the right panels. (B) DIC images showed the dorsal aorta
730
morphological integrity (arrows) in 24 hpi and control larvae. Note that blood cells were
731
visible inside dorsal aorta. (C) TUNEL assay and acridine orange staining were
732
performed in 24 hpi and control larvae to detect cells with DNA damage and necrotic
733
dead cells, respectively. Fli1a-GFP fluorescence is absent in infected fish, but TUNEL
734
positive cells were detected in the ventral part of the tail, which contains the caudal
735
hematopoietic tissue. No differences were found between control and SVCV-infected
736
larvae stained with acridine orange. (D) Fluorescent images of Fli1a-GFP of 3dpf
737
SVCV-infected and control larvae showing the loss of fluorescence in endothelial cells
738
after 22 hours of infection. Higher magnification images showed the intact morphology
739
of the dorsal aorta in infected larvae even when the fluorescence was lost (arrow head).
740
(E) The localization of L-plastin positive cells in caudal hematopoietic tissue is
741
consistent with that of TUNEL positive cells at 24 hours post-infection. TUNEL assay
742
was performed in 24 hours SVCV-infected and control larvae to detect cells with DNA
743
damage. L-plastin-positive cells are absent in infected fish, but TUNEL positive cells
744
are detected in the ventral part of the tail, which contains the caudal hematopoietic
745
tissue. (F) Expression qPCR measurements of vascular endothelium-related genes. The
746
samples were obtained from infected larvae at different times post-infection. Each
35
747
sample (technical triplicates) was normalized to the 18S gene. To avoid differences due
748
to the larval development stage, the samples were standardized with respect to an
749
uninfected control of the same age. Three biological replicates were used to calculate
750
the means ± standard error of mean. Significant differences between infected and non-
751
infected larvae at the same time point were displayed as *** (0.0001<p<0.001), **
752
(0.001<p<0.01) or * (0.01<p<0.05).
753
Figure 3. Viral detection and antiviral response measure over time. (A) Expression
754
qPCR measurements of TLRs and RLRs associated with viral detection. (B) Expression
755
qPCR measurements of antiviral response-related genes. Samples were obtained from
756
infected larvae at different times post-infection. Each sample (technical triplicates) was
757
normalized to the 18S gene. To avoid differences due to the larval development stage,
758
the samples were standardized with respect to an uninfected control of the same age.
759
Three biological replicates were used to calculate the means ± standard error of mean.
760
Significant differences were displayed as *** (0.0001<p<0.001), ** (0.001<p<0.01) or
761
* (0.01<p<0.05).
762
Figure 4. SVCV positive cells are distinct from endothelial cells. (A) Confocal
763
images (maximal projections from multiple z-stacks) of 3 dpf infected Tg(Fli1a-GFP)
764
larvae immunostained with SVCV antibody at 20 hpi (60x magnification). (B) (C) (D)
765
3D reconstructions of the different zones shown in A. Sequential steps of volume
766
clipping of the Fli1a channel are shown from left to right. We classified SVCV-positive
767
cells as circulating- (white arrows) and interacting- (white head arrows) cells. The axis
768
numbers correspond to μm.
769
Figure 5. SVCV positive cells are L-plastin positive and Fli1a negative. (A)
770
Confocal images 3D reconstruction of different scenes observed during the infection of
36
771
3 dpf larvae. Double positive cells are observed around vessels, interacting or not with
772
the endothelium. (B) Confocal images (maximal projections from multiple z-stacks) of
773
infected Tg(Fli1a-GFP) larvae double-immunostained with SVCV and L-plastin
774
antibodies at 20 hpi (60x magnification). Panels on the right show 3D reconstructions of
775
the areas identified in the central panels.
776
Figure 6. Neutrophils and macrophages respond to SVCV. (A) Confocal images
777
(maximal projections from multiple z-stacks) of infected 3 dpf Tg(Mpx-GFP) larvae
778
immunostained with L-plastin antibody 24 hpi (20x magnification). Head showing the
779
loss of L-plastin-positive cells in infected fish. MPX-positive cells increase in the same
780
area as a consequence of the migratory wave from the yolk sac population. (B) View of
781
yolk sac and tail of infected and control larvae. A yolk sac neutrophil population was
782
observed in the control larvae. Macrophages are missing in infected larvae. (C)
783
Neutrophil count in the head of control and 24 hours SVCV-infected larvae (n=4). (D)
784
qPCR results of the dynamics of macrophages and neutrophils population during SVCV
785
infection. (E) PU.1 MO was used to block macrophages development. Macrophages-
786
depleted fish showed a delay in mortality of 9 hours after SVCV infection at 2 dpf
787
compared to controls. The results are representative of 3 independent experiments of 2
788
or 3 biological replicates each. qPCR of marco and mpx facilitated an assessment of the
789
efficiency of PU.1 MO at 48 hpi. Regarding SVCV N gene, the morphants showed less
790
viral transcription during first hours of infection, suggesting that macrophages are the
791
main target of SVCV during the first hours of infection. Significant differences were
792
displayed as *** (0.0001<p<0.001), ** (0.001<p<0.01) or * (0.01<p<0.05).
793
Figure 7.SVCV elicits an inflammatory response and also macrophages pyroptosis.
794
(A) Expression qPCR measurements of some important inflammatory response genes.
37
795
Three biological replicates were used to calculate the means ± standard error of mean.
796
Significant differences between infected and non-infected larvae at the same time point
797
were displayed as *** (0.0001<p<0.001), ** (0.001<p<0.01) or * (0.01<p<0.05). (B)
798
Confocal images (maximal projections from multiple z-stacks) of 3 dpf infected
799
Tg(Mpeg1-GFP) larvae immunostained with SVCV and Caspa antibodies at 22 hpi (60x
800
magnification). The last macrophages in the larvae were Caspa positive. Some of these
801
cells were infected (boxed area) and the other cells were interacting with a group of
802
infected cells (arrowhead). (The arrow shows potential macrophage pyroptosis,
803
observed as visible membrane vesicle shedding. The last panel shows a 3D
804
reconstruction of the infected macrophage in the boxed area, with a viral antigen
805
positive area on its surface (arrowhead). (C) Resting macrophages shown Caspa signal
806
inside the cell. The level of Caspa was significantly less in resting macrophages than
807
that in the macrophages of infected larvae. (D) Confocal images (maximal projections
808
from multiple z-stacks) of infected Tg(Mpeg1-GFP) larvae immunostained with SVCV
809
and Il1 antibodies at 20 hpi (60x magnification). Regarding the proportion of Il1 and
810
SVCV fluorescence observed, we distinguished two types of macrophages. (E) 3D
811
reconstruction of infected macrophages in the boxed area is displayed. Il1 signal is
812
visible inside cells, as the SVCV signal. SVCV antigens are also visible on the cell
813
surface. (F) 3D reconstruction of non-infected macrophages positive for Il1 in boxed is
814
displayed. Mpeg1 fluorescence is lost in favor of the Il1 signal. Cellular debris positive
815
for SVCV is visible around these macrophages.
816
Figure 8. Interleukin 1 release is induced by SVCV infection. (A) Resting
817
macrophages shown Il1 signal inside the cell. The level of il1 was significantly less
818
in resting macrophages than that in the macrophages of infected larvae. (B) The
38
819
macrophages in 3 dpf infected larvae showed shedding of Il1-positive vesicles. (C)
820
Il1 also could be released directly through membrane pores in the macrophages of
821
infected larvae. (D) In vivo PI staining of infected larvae showing macrophages PI
822
uptake through membrane pores. (E) 3D reconstruction of macrophages in an SVCV
823
infected larvae positive for IL1β and TUNEL are displayed.
39
Download