Download Mutational analysis of the glycosaminoglycan

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Henipavirus wikipedia , lookup

Influenza A virus wikipedia , lookup

Transcript
1
A sequence of basic residues in the porcine circovirus type 2 capsid protein
2
is crucial for its co-expression and co-localization with the replication
3
protein
4
Liping Huanga,b,+, Nicolaas Van Rennea,+, Changming Liub, Hans J. Nauwyncka,*
5
6
Laboratory of Virology, Faculty of Veterinary Medicine, Ghent University, Salisburylaan
7
133, B-9820 Merelbeke, Belgiuma; Division of Swine Infectious Diseases, State Key
8
Laboratory of Veterinary Biotechnology, Harbin Veterinary Research Institute, The Chinese
9
Academy of Agricultural Sciences, Maduan Street 427, Harbin 150001, Chinab
10
11
Category: Animal: DNA viruses
12
13
Running Title: Function analysis of 97RIRKVK102 in PCV2 Capsid Protein
14
15
+
L.H. and N.V.R. contributed equally to this work
16
17
18
*
Address correspondence to Hans J. Nauwynck, [email protected]
Tel.: +32 9 264 73 73
19
20
Number of words in Abstract: 149
21
Number of words in Text: 3,862
22
Number of Figures: 6
23
Number of Table: 0
1
24
Abstract
25
Porcine circovirus type 2 (PCV2) encodes two major proteins: the replication protein (Rep)
26
and the capsid protein (Cap). The capsid protein displays a conserved stretch of basic
27
residues situated on the inside of the capsid which role is hitherto unknown. We have used a
28
reverse-genetics approach to investigate its function and found that mutations in these amino
29
acids hindered Cap mRNA translation and hampered Cap-Rep co-localization, yielding unfit
30
viruses. Intriguingly, co-transfection with a wild type PCV2 virus of a different genotype
31
partially rescued mutant Cap expression, showing the importance of this basic pattern for
32
efficient translation of Cap mRNA into protein. Our work shows that Cap and Rep are
33
expressed independently of each other, and that this amino acid sequence of the Cap is vital
34
for viral propagation. This study provides a method for studying unfit PCV2 virions and
35
offers new insights into the intracellular modus vivendi of PCV2.
36
37
38
39
40
41
2
42
Introduction
43
To date, two types of porcine circovirus have been described, porcine circovirus type 1 and
44
2 (PCV1 and PCV2) (Allan et al., 1998; Tischer et al., 1982). PCV1 was originally reported
45
as a contaminant of the PK-15 cell line and is considered pathogenic only in young porcine
46
fetuses (Saha et al., 2011). PCV2 is associated with several clinical manifestations
47
collectively known as porcine circovirus diseases (PCVD) with postweaning multisystemic
48
wasting syndrome (PMWS) as the most significant manifestation (Segalés, 2012).
49
PCV2 belongs to the genus Circovirus of the family Circoviridae, of which four genotypes
50
have been described so far: PCV2a, PCV2b, PCV2c and PCV2d (Xiao et al., 2015). PCV2
51
is an extremely small non-enveloped virus with an icosahedral capsid measuring only 17 nm
52
in diameter. Its genome comprises a circular single-stranded DNA of only 1766–1769
53
nucleotides (Huang et al., 2011b) which encodes two major open reading frames (ORFs):
54
ORF1 and ORF2. ORF1 encodes the non-structural replication protein Rep and its splice
55
variant Rep', which are both necessary for viral replication (Mankertz and Hillenbrand, 2001;
56
Mankertz et al., 1998a). ORF2 is translated into the capsid protein (Cap) (Nawagitgul et al.,
57
2000), 60 copies of which make up the icosahedral proteinaceous shell protecting the viral
58
genome (Crowther et al., 2003). Cap is the primary immunogenic protein, which has the
59
ability to bind to the host cell receptor (Misinzo et al., 2006).
3
60
As PCV2 exploits heparin sulfate or dermatan sulfate side chains of cell surface
61
proteoglycans as attachment receptors, the putative heparin-binding motif 98IRKVKV103 was
62
initially thought to be a potential binding site (Misinzo et al., 2006). Recently however, the
63
atomic structure of PCV2 Cap revealed this motif was located on the inside of the capsid,
64
excluding a role in cell-surface attachment (Khayat et al., 2011). Moreover, it appeared to be
65
part of a larger constellation of positively charged amino acids on the inner surface of the
66
capsid, including a notable all-conserved 97RIRKVK102 sequence strongly reminiscent of the
67
putative heparin-binding motif. The manifest conservation of this pattern suggests a
68
well-defined purpose for this specific stretch of basic residues. The present study employs a
69
reverse-genetics approach to clarify its function by mutational analysis.
70
71
Results
72
Mutations in the basic residues of 97RIRKVK102 decrease Cap expression and interfere
73
with subcellular Cap localization
74
In order to determine the role of the basic amino acids in the conserved
75
sequence,
76
M3(R97T,R99T,K100T,K102T) were constructed. Sequencing confirmed that all mutations
77
were introduced correctly. PCV-negative PK-15 cells were transfected with DNA of each
mutants
M1(R97T,R99T),
4
97
RIRKVK102
M2(K100T,K102T)
and
78
mutant and the parental clone, and subsequently immunostained at different time points
79
after transfection. The numbers of PCV2-positive cells are shown in Fig. 1a. For
80
M1(R97T,R99T), M2(K100T,K102T) and the parental clone, Caps were detected starting
81
from 12 hours post-transfection (hpt), and the numbers of positive cells rapidly increased to
82
a maximum at 36 hpt. For M3(R97T,R99T,K100T,K102T) however, not a single
83
Cap-positive cell was ever detected. In the first 36 hours post-transfection, the number of
84
PCV2-positive cells was comparable between M1(R97T,R99T), M2(K100T,K102T) and
85
the parental clone. However at 72 hpt, the number of positive cells markedly decreased for
86
these mutant viruses, suggesting a drop in viral fitness in secondary rounds of replication.
87
Despite similar numbers of positive cells at 36 hpt, immunostaining showed a strikingly
88
different Cap expression-pattern (Fig. 1b). Cap staining intensity of M1(R97T;R99T) was
89
weaker than the parental clone, and signals for M2(K100T,K102T) were even less than
90
those of M1(R97T,R99T). M3(R97T,R99T,K100T,K102T) signal was completely absent.
91
This prompted the question if the introduced mutations caused conformational changes
92
hindering antibody binding, or whether the protein itself was expressed at a lower level.
93
Western blotting of transfected PK-15 cell lysates confirmed there was a reduction in the
94
amount of Cap for M1(R97T,R99T) and M2(K100T,K102T), and a complete absence of
95
Cap for M3(R97T,R99T,K100T,K102T) (Fig. 2a). On the contrary, Rep immunostaining
5
96
never displayed any difference between wild type and mutants, not even for
97
M3(R97T,R99T,K100T,K102T) whose Cap expression was completely abolished
98
3a). Taken together, these results demonstrate that the mutants were only hampered in Cap
99
but not Rep expression, and that Cap expression occurs independently of Rep.
(Fig.
100
In addition to expression alterations, a difference in subcellular distribution was observed
101
for Cap among the different mutants and the parental clone at 36 hpt (Fig. 1b). Of the
102
parental clone-transfected cells, 3.2% showed Cap presence in both nucleus and cytoplasm,
103
and another 5.1% had a positive Cap signal in the nucleus only. For M1(R97T,R99T) and
104
M2(K100T,K102T), the number of cells that showed Cap in both nucleus and cytoplasm
105
dropped dramatically (1.2 and 0.8% respectively), but this decrease was completely
106
compensated by an increase in Cap presence in the nucleus only (8.2% and 8.8%). Cap was
107
never observed in the cytoplasm only. This finding shows that mutations in the
108
97
109
rather than maintaining a cytoplasmic stage in its lifecycle.
110
To determine whether the decrease in Cap expression in the cells coincided with a decrease
111
of viral capsids released from the transfected PK-15 cells, capture-ELISA kinetics of
112
transfected cell supernatants were studied (Fig. 2b). Throughout a time course of 72h,
113
OD450nm values of M1(R97T,R99T) and M2(K100T,K102T) remained lower than the
RIRKVK102 pattern result in Cap being shunted towards the nucleus of transfected cells,
6
114
parental clone. As expected, OD450nm values for M3(R97T,R99T;K100T,K102T) were
115
negative. Moreover, the data are perfectly in line with the intracellular capsid production,
116
and indicate that the conserved stretch of basic residues has no role in viral egress once the
117
virions are assembled.
118
119
Mutations in the 97RIRKVK102 hamper Cap-Rep co-localization
120
To determine the potential co-localization between Cap and Rep, PK-15 cells were
121
transfected at 24 hpt and stained with Cap-specific mAb 12E12 (Red) and Rep-specific
122
mAb F210 D6 (Green). M1(R97T,R99T) and especially M2(K100T,K102T) displayed a
123
striking lack of Cap-Rep co-localization (Fig. 3a). The co-localization events were
124
quantified by CoLocolizer Pro 2.7.1 using Manders’ coefficient correlation. This analysis
125
revealed a Cap-Rep co-localization rate of 81%, 56% and 24% for the parental clone,
126
M1(R97T,R99T) and M2(K100T,K102T) respectively. For M3(R97T,R99T,K100T,K102T)
127
a co-localization coefficient could not be calculated since Cap was not expressed (Fig. 3b).
128
129
Mutations in the 97RIRKVK102 drastically impact viral fitness
130
To investigate whether the mutant capsids released in the supernatant were still infectious,
131
we determined the viral titers of the supernatants collected at different time points after
7
132
transfection (Fig. 4a). A time-dependent rise of viral titers was observed starting from 24
133
hpt for M1(R97T,R99T), M2(K100T,K102T) and the parental clone. The highest titers were
134
found at 60 hpt: 3.9 log10TCID50/ml for M1(R97T,R99T), 3.3 log10TCID50/ml for
135
M2(K100T,K102T) and 5.3 log10TCID50/ml for the parental clone. The titers of the mutants
136
were lower than that of parental clone, reflecting the dramatic decrease in Cap expression
137
observed by immunostaining, western blot and ELISA. M3(R97T,R99T,K100T,K102T)
138
supernatants were always negative.
139
In order to test the fitness of the progeny virus, PCV2-transfected cell supernatants were
140
passaged on PCV-negative PK-15 cells. A gradual decrease in viral titers reaching a level of
141
non-detection was observed from passages two to seven in case of the mutants, while a
142
gradual increase in titers was observed for the parental clone (Fig. 4b). PCV2 genomic
143
sequences obtained from passage two, five and seven for parental virus, and passage two
144
and four for M1(R97T,R99T) and passage two for M2(K100T,K102T) were all 100%
145
identical to the genome in the pCR-BluntII-TOPO vector.
146
Taken together, this shows that Caps with mutations in the conserved basic sequence can
147
reach the assembly stage and properly exit the cell for secondary rounds of infection.
148
However, the impact on protein expression is so enormous that overall viral fitness is
149
seriously compromised. Without this pattern intact, the virus is doomed to disappear from the
8
150
gene pool, recapitulating the selection pressure observed in vivo.
151
152
The decrease in mutant Cap expression is exclusively due to an altered amino acid
153
sequence which results in a decrease in Cap mRNA translation
154
Next, we quantified Cap mRNA using RT-qPCR to see if the decrease in Cap expression of
155
the mutant viruses was caused by a decrease in Cap mRNA transcription. All samples
156
derived from the transfected PK-15 cells were positive at 12 hpt (Fig. 5a). Notwithstanding
157
the differences in Cap mRNA at 36 hpt of the mutants compared to the parental clone, the
158
decrease was not strong enough to fully explain the drastic decrease or even absence of
159
protein expression. At 72 hpt, which corresponds with the next replication cycle, Cap
160
mRNA copy number was approximately 80% lower than the parental clone, which can be
161
explained by the reduced viral fitness of the mutants in subsequent replication cycles. This
162
clearly shows that the dramatic reduction in Cap protein expression at 36 hpt is not solely
163
due to a decrease in Cap mRNA.
164
In order to completely exclude that nucleotide changes in the viral DNA or the Cap mRNA
165
would interfere with translation processes inside the cell, a new synonymous mutant, M4,
166
was constructed. M4 has four silent nucleotide changes, which do not alter the amino acid
167
sequence of Cap. We hypothesized that, if changes on the nucleotide level would influence
9
168
transcription or translation, M4 and M3(R97T,R99T,K100T,K102T) fitness would be evenly
169
affected, since both have four nucleotide substitutions in exactly the same region.
170
Interestingly, M4 displayed an exactly similar phenotype as the parental clone, yielding
171
infectious, fit viruses (Fig. 5b and 5c). This excludes an effect of the nucleotide sequence in
172
itself, and confirms that the dramatic effects on viral viability are entirely due to the altered
173
amino acids in the mutants.
174
The inability of mutant Caps to be properly expressed in spite of ample mRNA presence
175
could be explained by proteasomal degradation of newly expressed Cap proteins. However,
176
PK-15 cells treated with proteasome inhibitor MG132 for 2h at 24hpt did not rescue proper
177
Cap expression (data not shown). The picture that emerges is that mutant Cap mRNA
178
transcripts are hampered in translation, rather than Cap proteins being degraded once
179
expressed.
180
181
Co-transfection with a fit genotype PCV2a virus partially rescues mutant Cap
182
translation through an unknown mechanism
183
Our findings led us to believe that the conserved basic sequence was critical in getting Cap
184
mRNA efficiently translated into Cap. If so, co-transfection of the 48285-based mutants
185
(genotype PCV2b) with the PCV2a-LG strain should rescue mutant Cap expression as a
10
97
RIRKVK102 pattern. First, we verified that mouse mAb 8G12 only
186
result of its intact
187
reacted to PCV2a-LG Cap and did not cross-react with PCV2b mutants. Similarly, mouse
188
mAb 6E9 only detected 48285-based mutant capsids, and did not cross react with
189
PCV2a-LG Cap (data not shown). By staining mutant-transfected cells with mAb 6E9 we
190
obtained identical results as with mAb 38C1 (Fig. 6 left column). As expected,
191
co-transfection with PCV2a-LG did not have any effect on LG expression (Fig. 6 right
192
column). Intriguingly, this procedure partially rescued mutant Cap translation, which was
193
particularly visible in M3(R97T,R99T,K100T,K102T)-LG co-transfection (Fig. 6 middle
194
column). This highlights once more the importance of an intact 97RIRKVK102 alignment for
195
efficient Cap translation.
196
197
Discussion
198
Being the smallest of all known mammal viruses with limited genomic coding capacity,
199
porcine circoviruses are truly fascinating biological systems. Despite their reliance on
200
proofreading-competent host polymerases for replication, their genomes mutate
201
exceptionally rapidly while keeping an extremely low genetic diversity within the different
202
PCV2 genotypes (Firth et al., 2009). In other words, most nucleotide substitutions
203
disappear immediately under intense purifying selection, and as a consequence, PCV2
11
204
genomes are allowed only marginal leeway to evolve. This should not come as a surprise,
205
since every part of its genome is exploited on multiple levels: the stem-loop and adjacent
206
motifs need intense conservation for (i) binding and cleavage by the replication complex
207
(Steinfeldt et al., 2001), (ii) the Cap mRNA splicing acceptor site (Mankertz et al., 1998b)
208
and (iii) the larger part of the Rep promoter sequence (Mankertz and Hillenbrand, 2002).
209
The region clockwise of the stem-loop cannot be tampered with for its DNA encodes (i) the
210
replication protein Rep and its frame-shifted splice variant Rep’ (Mankertz and Hillenbrand,
211
2001) as well as (ii) ORF3 (Liu et al., 2005) and (iii) ORF4 (He et al., 2013) on the amino
212
acid level. On a nucleotide level, this stretch also contains sequences for (iv) the Rep’
213
mRNA splicing donor and acceptor sites (Mankertz and Hillenbrand, 2001), and (v and vi)
214
the Cap mRNA splicing donor site and the Cap promoter (Mankertz and Hillenbrand, 2002;
215
Mankertz et al., 1998b). The genomic part counterclockwise to the stem loop is supposed to
216
have more freedom for variation because it only encodes the Cap on amino acid level
217
(Nawagitgul et al., 2000), and the smaller part of the Rep promoter (Mankertz and
218
Hillenbrand, 2002) on the nucleotide level. This understanding makes any conserved
219
sequence of ORF2 an interesting target for investigation.
220
Since cryo-EM reconstruction of PCV2 VLPs demonstrated that the all-conserved
221
98
IRKVKV103 motif originally identified by Misinzo (2006) was situated on the inside of the
12
222
capsid, and took part in a larger constellation of positively charged amino acids comprising
223
the basic sequence
224
negatively charged viral ssDNA strand upon encapsidation. That this stretch of basic residues
225
was also implicated in Cap mRNA translation (Fig.2 and Fig.5) came as a complete surprise
226
and the reason for it still mystifies us. It is hard to craft an explanation without falling into a
227
chicken or egg causality dilemma. One might think that Cap mRNA needs the Cap to be
228
efficiently exported from the nucleus, but for Cap to come into existence an efficient viral
229
mRNA export would be required first. Likewise, the Cap might act as a molecular switch
230
turning on Cap translation machinery, but this would require prior translation of Cap! We
231
tried to solve this catch-22 situation by co-transfecting with a wild typePCV2 strain, but even
232
then we see a preferential translation of wild-type PCV2 (Fig. 6). This argues in favor of a
233
vision where the basic amino acid sequence is not so much a biological switch enhancing
234
translation, but rather takes part in a molecular interface interacting with another factor. One
235
can imagine that the basic amino acids of the capsid protein have to be shielded from
236
aspecific interactions with other proteins and nucleic acids, before being imported to the
237
nucleus where they are deposited on viral DNA with Swiss-watch precision. In many ways,
238
they resemble histones safeguarded and escorted by histone chaperones. Interestingly, an
239
interaction between Cap and histone chaperones NAP1 and NPM has been described
97
RIRKVK102, we understood these basic amino acids stabilize the
13
240
(Finsterbusch et al., 2009). It is tempting to think that the removal of the positively charged
241
amino
242
M3(R97T,R99T,K100T,K102T) Caps abolishes their interaction with co-translational
243
chaperones, causing any further Cap mRNA translation to grind to a halt and leaving nascent
244
Cap polypeptides stuck on the negatively charged ribosome. In this respect, the partial
245
rescue of mutant Cap translation we see by co-transfecting with a wild type PCV2a virus (Fig.
246
6) may be caused by a mutant Cap-wild type Cap interaction which releases the mutant Cap
247
from the ribosome, and allows translation to continue, albeit still inefficiently. PCV1 capsid
248
proteins display a virtually identical 98RIRKAK103 sequence, and it would be interesting to
249
see if M3(R97T,R99T,K100T,K102T) would still be rescued after co-transfection with a
250
PCV1 Cap expression construct. PCV1 Caps are expected to have inferior binding
251
capabilities with PCV2 Caps because of their antigenic differences in the capsid protein.
252
PCV1 should therefore not be able to restore M3(R97T,R99T,K100T,K102T) expression if
253
rescue is indeed dependent on an interaction with wild type Cap. In addition, an unbiased
254
siRNA screen targeting histone chaperones could identify which chaperones, if any, are
255
involved in PCV2 Cap translation.
256
As a final remark, it is conceivable that the replication protein itself exerts some additional
257
chaperone activity. Rep may just guide Cap subunits to the right sub-nuclear compartment
acid
pattern
in
M1(R97T,R99T),
14
M2(K100T,K102T)
and
258
for further assembly, or even more speculative, Rep might simply drive genome
259
encapsidation. As a matter of fact, all circovirus replication proteins share a common
260
superfamily 3 helicase domain (Roux et al., 2013) which assemble into ring-shaped
261
multimeres responsible for unwinding dsDNA by means of an ATPase-powered molecular
262
motor (Hickman and Dyda, 2005). Should this activity be performed in close proximity to
263
capsid proteins, it is not improbable that Rep subunits form replicative complexes displacing
264
the viral strand towards adjacent capsid subunits, thus facilitating the process of genome
265
packaging in complete harmony with replication. Hypothetical as this may be, also in the
266
closely related beak and feather disease virus (BFDV), an elegant relationship exists between
267
the two viral proteins, for Rep nuclear localization depends entirely on a cytoplasmic
268
interaction with Cap, after which both shuttle to the nucleus (Heath et al., 2006). These
269
findings all support a vision for circoviruses in which an intimate relationship exists between
270
Cap and Rep which has been underestimated until now, and can serve as a working model for
271
future experimental studies.
272
In conclusion, this study shows that the conserved stretch of basic residues
273
comprising 97RIRKVK102 of the PCV2 Cap is crucial for its expression. Mutations in this
274
pattern resulted in (i) a reduced translation of Cap, (ii) hindered Cap-Rep
275
co-localization and (iii) finally unfit PCV2 virions. Additionally, co-transfection with a
15
276
wild type PCV2 virus partially rescues mutant Cap expression, highlighting the importance
277
of these amino acids for an efficient translation of Cap mRNA into protein. We speculate
278
that 97RIRKVK102 is part of a molecular interface interacting with co-translational
279
chaperones, however, future experiments need to be carried out to confirm this hypothesis.
280
281
282
Material and methods
283
Virus, cells and virus titration
284
PCV2 strain 48285 (Accession number: AF055394) clone 48285-24, PK-15 cells and
285
culture conditions were described earlier (Saha et al., 2012b). PCV2 strain LG (Accession
286
number: HM038034) clone pMD18/PCV2a-LG was described earlier (Huang et al., 2011b).
287
Virus titration was performed as previously described (Meerts et al., 2005).
288
289
Monoclonal antibodies against PCV2
290
Cap-binding mouse monoclonal antibodies (mAb) 38C1, 12E12 and 6E9 were generated in
291
our laboratory and described earlier (Saha et al., 2012a). MAb 8G12 was generated against
292
PCV2a-LG with a procedure described by Huang et al. (2011a). Rep-binding mAb F210
293
D6 was described previously (McNeilly et al., 2001). For western blotting, mAb 2E8 was
16
294
generated against PCV2-Cap derived from PCV2a strain LG (Huang et al., 2011a) and is
295
directed against the linear epitope 26–36aa of PCV2-Cap (Guo et al., 2011).
296
297
Construction of mutants
298
A site-directed mutagenesis was performed by targeting the basic amino acids of
299
97
300
by an uncharged hydrophilic threonine (T) using a Quickchange® site-directed mutagenesis
301
kit (Stratagene, La Jolla, USA) based on clone 48285-24. R at positions 97 and 99 were
302
substituted by T using primers M1-FW (5’-atactacacaataacaaaggttaaggttgaat-3’) and
303
M1-REV (5’-attcaaccttaacctttgttattgtgtagtat-3’) resulting in the mutant M1(R97T,R99T).
304
Mutant M2(K100T,K102T) was constructed by mutating K into T at positions 100 and 102
305
using
306
(5’-attcaaccgtaaccgttcttattctgtagtat-3’).
307
constructed by replacing R or K with T at all four positions 97, 99, 100 and 102 using
308
primers
309
(5’-attcaaccgtaaccgttgttattgtgtagtat-3’). Mutant M4 was constructed by changing the codons
310
(aga ata aga aag gtt aag) of
311
M4-FW
RIRKVK102 in the Cap of PCV2 strain 48285. Arginine (R) or lysine (K) was substituted
primers
M2-FW
M3-FW
(5’-atactacagaataagaacggttacggttgaat-3’)
Mutant
M2-REV
M3(R97T,R99T,K100T,K102T)
(5’-atactacacaataacaacggttacggttgaat-3’)
97
and
and
was
M3-REV
RIRKVK102 into (agg ata agg aaa gtt aaa) using primers
(5’-atactacaggataaggaaagttaaagttgaat-3’)
17
and
M4-REV
(5’-
312
attcaactttaactttccttatcctgtagtat-3’), resulting in a synonymous mutant with unaltered amino
313
acid sequence.
314
315
Transfections
316
Parental and mutated plasmids were digested with Hind Ⅲ and subsequently purified
317
genomic DNA fragments were self-ligated for 20 mins at room temperature using T4 DNA
318
ligase (Invitrogen, Merelbeke, Belgium). 800 ng of DNA was then transfected onto
319
PCV-negative PK-15 cells with lipofectamine 2000 (Invitrogen, Merelbeke, Belgium)
320
according to the manufacturer's instructions. Empty-plasmid transfected PK-15 cells were
321
included as a mock.
322
323
Immunoflurescence assays and co-localization quantification
324
Methanol-fixed transfected cells were stained with a previously described indirect
325
immunofluorescence assay (Saha et al., 2010). MAb 38C1 and fluorescein isothiocyanate
326
(FITC) labeled goat anti-mouse IgG (Molecular Probes, Eugene, USA) were used as the
327
primary and secondary antibodies, respectively. Images were made with Leica TCS SP2
328
confocal microscope (Leica Microsystems GmbH, Mannheim, Germany). For double
329
immunofluorescence staining for Cap and Rep, transfected PK-15 cells were first incubated
18
330
with anti-Cap mAb 12E12 (1:2) and then with Alexa Fluor 594 goat-anti mouse IgG2b
331
(Molecular Probes; 1:200). Afterwards, the cells were stained for PCV2 Rep by incubation
332
with anti-Rep mAb F210 D6 (1:500). Subsequently, a 1:200 dilution of FITC-labeled
333
goat-anti mouse IgG1 (Molecular Probes) was added. All antibodies were incubated for 1 h
334
at 37°C. The cells were washed three times with PBS after incubation with primary and
335
secondary antibodies. Specificity of the staining for different protein was demonstrated
336
using an irrelevant, isotype-matched mAb II1H/4-5G (Costers et al., 2010) and 13D12
337
(Nauwynck and Pensaert, 1995) and by the complete absence of Cap- and Rep-specific
338
fluorescence in mock-transfected PK-15 cells. The co-localization of Cap and Rep was
339
analyzed using the software CoLocalizer Pro 2.7.1 as described previously (Zinchuk and
340
Zinchuk, 2008). Background correction of the images was performed prior to coefficient
341
calculation. The co-localization rates were measured based on Manders’ coefficient, which
342
varies from 0 to 1. A coefficient value of 0 to 0.6 indicates absence of co-localization while
343
a value of 0.6 to 1.0 indicates co-localization.
344
345
Western blot analysis
346
Western-blot analysis was used to quantify the amount of Cap produced by the different
347
mutants
and
the
parental
clone.
Briefly,
19
M1(R97T,R99T),
M2(K100T,K102T),
348
M3(R97T,R99T,K100T,K102T), parental clone and mock-transfected PK-15 cells were
349
harvested at 36 hpt. Sodium dodecyl sulfate polyacrylamide gel electrophoresis
350
(SDS-PAGE) and Western-blot were performed as described previously (Lefebvre et al.,
351
2008). Anti-PCV2-Cap mAb 2E8 (Guo et al., 2011; Huang et al., 2011a), a
352
PCV2-irrelevant mAb 1C11 (Nauwynck and Pensaert, 1995), HRP-conjugated α-tubulin
353
(Cell signaling technology, Danvers, MA, USA) and HRP-conjugated sheep anti-mouse
354
IgG (H+L) (Invitrogen) were used as the primary antibody, negative control, loading
355
control and secondary antibody, respectively. Antigen-antibody complexes were visualized
356
using an enhanced chemiluminescence assay (Amersham Biosciences, NJ, USA).
357
358
Capture-ELISA
359
To quantify PCV2-Cap in the supernatants collected at different time points after
360
transfection, a capture-ELISA was performed as previously described (Huang et al., 2011a)
361
with minor modifications. Purified mAb 38C1 (5 μg/ml), biotin-labeled 38C1 (1:1000) and
362
Streptavidin-biotinylated horse-radish peroxidase (HRP) complex (1:1000; GE Healthcare,
363
United Kingdom) were used as coating, detecting and secondary antibody, respectively. The
364
color was developed and the reaction was stopped as described (Saha et al., 2012a). The
365
optical density (OD) at 450 nm was measured by a microplate reader (Bio-Rad, Hercules,
20
366
CA, USA). The capture-ELISA was performed in triplicate. The mean OD450 nm value of
367
the mock plus three standard deviations were calculated to use as a cut-off value of the
368
capture-ELISA.
369
370
RT-qPCR
371
Total RNA was extracted from transfected cells using the RNeasy Mini Kit (Qiagen, Hilden,
372
Germany). For each sample, 500 ng of RNA was reverse transcribed into cDNA using
373
SuperScriptTM First-Strand Synthesis System (Invitrogen, Merelbeke, Belgium).
374
The
375
(5’-gctccacactccatcagtatgac-3’) used for quantifying Cap mRNA transcript were adapted
376
from (Yu et al., 2007) with modifications based on sequence data of PCV2 strain 48285.
377
The reverse primer was designed to span the splice junction to ensure that PCV2 DNA was
378
not amplified. qPCR was performed with Applied Biosystems StepOnePlusTM Real-Time
379
PCR system (Genetic Systems Company, Foster City, California) using SensiMixTM SYBR
380
mastermix (Bioline, London, UK).
forward
primer
(5’-agatgccatttttccttctc-3’)
and
reverse
primer
381
382
Acknowledgements
383
We would like to thank Carine Boone and Lieve Sys for their excellent technical assistance.
21
384
This study was supported by the National Natural Science Foundation of China (Grant No.
385
31302110).
22
386
References
387
Allan, G. M., McNeilly, F., Kennedy, S., Daft, B., Clarke, E. G., Ellis, J. A., Haines, D. M., Meehan, B. M.
388
& Adair, B. M. (1998). Isolation of porcine circovirus-like viruses from pigs with a wasting disease
389
in the USA and Europe. Journal of veterinary diagnostic investigation 10, 3-10.
390
Costers, S., Lefebvre, D. J., Van Doorsselaere, J., Vanhee, M., Delputte, P. L. & Nauwynck, H.J. (2010).
391
GP4 of porcine reproductive and respiratory syndrome virus contains a neutralizing epitope that is
392
susceptible to immunoselection in vitro. Archive of virology 155, 371-378.
393
Crowther, R. A., Berriman, J. A., Curran, W. L., Allan, G. M. & Todd, D. (2003). Comparison of the
394
structures of three circoviruses: chicken anemia virus, porcine circovirus type 2, and beak and
395
feather disease virus. Journal of virology 77, 13036-13041.
396
Finsterbusch, T., Steinfeldt, T., Doberstein, K., Rodner, C. & Mankertz, A. (2009). Interaction of the
397
replication proteins and the capsid protein of porcine circovirus type 1 and 2 with host proteins.
398
Virology 386, 122-131.
399
Firth, C., Charleston, M. A., Duffy, S., Shapiro, B. & Holmes, E. C. (2009). Insights into the evolutionary
400
history of an emerging livestock pathogen: porcine circovirus 2. Journal of virology 83,
401
12813-12821.
402
403
404
405
Guo, L., Lu, Y., Huang, L., Wei, Y. & Liu, C. (2011). Identification of a new antigen epitope in the nuclear
localization signal region of porcine circovirus type 2 capsid protein. Intervirology 54, 156-163.
He, J., Cao, J., Zhou, N., Jin, Y., Wu, J. & Zhou, J. (2013). Identification and functional analysis of the
novel ORF4 protein encoded by porcine circovirus type 2. Journal of virology 87, 1420-1429.
406
Heath, L., Williamson, A. L. & Rybicki, E. P. (2006). The capsid protein of beak and feather disease virus
407
binds to the viral DNA and is responsible for transporting the replication-associated protein into the
408
nucleus. Journal of virology 80, 7219-7225.
409
410
Hickman, A.B. & Dyda, F. (2005). Binding and unwinding: SF3 viral helicases. Current opinion in
structural biology 15, 77-85.
411
Huang, L., Lu, Y., Wei, Y., Guo, L. & Liu, C. (2011a). Development of a blocking ELISA for detection of
412
serum neutralizing antibodies against porcine circovirus type 2. Journal of virological methods 171,
413
26-33.
414
Huang, L. P., Lu, Y. H., Wei, Y. W., Guo, L. J. & Liu, C. M. (2011b). Identification of one critical amino
415
acid that determines a conformational neutralizing epitope in the capsid protein of porcine circovirus
416
type 2. BMC Microbiology 11, 188.
417
418
Khayat, R., Brunn, N., Speir, J. A., Hardham, J. M., Ankenbauer, R. G., Schneemann, A. & Johnson, J.
E. (2011) The 2.3-angstrom structure of porcine circovirus 2. Journal of virology 85, 7856-7862.
419
Lefebvre, D. J., Costers, S., Van Doorsselaere, J., Misinzo, G., Delputte, P. L. & Nauwynck, H. J. (2008).
420
Antigenic differences among porcine circovirus type 2 strains, as demonstrated by the use of
23
421
monoclonal antibodies. Journal of general virology 89, 177-187.
422
Liu, J., Chen, I. & Kwang, J. (2005). Characterization of a previously unidentified viral protein in porcine
423
circovirus type 2-infected cells and its role in virus-induced apoptosis. Journal of virology 79,
424
8262-8274.
425
426
427
428
429
430
431
432
433
434
Liu, Q., Tikoo, S. K., Babiuk, L. A. (2001). Nuclear localization of the ORF2 protein encoded by porcine
circovirus type 2. Virology 285, 91-99.
Mankertz, A. & Hillenbrand, B. (2001). Replication of porcine circovirus type 1 requires two proteins
encoded by the viral rep gene. Virology 279, 429-438.
Mankertz, A. & Hillenbrand, B. (2002). Analysis of transcription of Porcine circovirus type 1. Journal of
general virology 83, 2743-2751.
Mankertz, A., Mankertz, J., Wolf, K. & Buhk, H. J. (1998a). Identification of a protein essential for
replication of porcine circovirus. Journal of general virology 79, 381-384.
Mankertz, J., Buhk, H. J., Blaess, G. & Mankertz, A. (1998b). Transcription analysis of porcine circovirus
(PCV). Virus Genes 16, 267-276.
435
McNeilly, F., McNair, I., Mackie, D. P., Meehan, B. M., Kennedy, S., Moffett, D., Ellis, J., Krakowka, S.
436
& Allan, G. M. (2001). Production, characterisation and applications of monoclonal antibodies to
437
porcine circovirus 2. Archive of virology 146, 909-922.
438
Misinzo, G., Delputte, P. L., Meerts, P., Lefebvre, D. J. & Nauwynck, H. J. (2006). Porcine circovirus 2
439
uses heparan sulfate and chondroitin sulfate B glycosaminoglycans as receptors for its attachment to
440
host cells. Journal of virology 80, 3487-3494.
441
Nauwynck, H. J. & Pensaert, M. B. (1995). Effect of specific antibodies on the cell-associated spread of
442
pseudorabies virus in monolayers of different cell types. Archive of virology 140, 1137-1146.
443
Nawagitgul, P., Morozov, I., Bolin, S. R., Harms, P. A., Sorden, S. D. & Paul, P. S. (2000). Open reading
444
frame 2 of porcine circovirus type 2 encodes a major capsid protein. Journal of general virology 81,
445
2281-2287.
446
Roux, S., Enault, F., Bronner, G., Vaulot, D., Forterre, P. & Krupovic, M. (2013). Chimeric viruses blur
447
the borders between the major groups of eukaryotic single-stranded DNA viruses. Nature
448
Communications 4, 2700.
449
Saha, D., Huang, L., Bussalleu, E., Lefebvre, D. J., Fort, M., Van Doorsselaere, J. & Nauwynck, H. J.
450
(2012a). Antigenic subtyping and epitopes' competition analysis of porcine circovirus type 2 using
451
monoclonal antibodies. Veterinary microbiology 157, 13-22.
452
Saha, D., Lefebvre, D. J., Ducatelle, R., Doorsselaere, J. V. & Nauwynck, H. J., (2011). Outcome of
453
experimental porcine circovirus type 1 infections in mid-gestational porcine foetuses. BMC
454
Veterinary research 7, 64.
455
Saha, D., Lefebvre, D. J., Ooms, K., Huang, L., Delputte, P. L., Van Doorsselaere, J. & Nauwynck, H. J.
456
(2012b). Single amino acid mutations in the capsid switch the neutralization phenotype of porcine
24
457
circovirus 2. Journal of general virology 93, 1548-1555.
458
Saha, D., Lefebvre, D. J., Van Doorsselaere, J., Atanasova, K., Barbe, F., Geldhof, M., Karniychuk, U. U.
459
& Nauwynck, H. J. (2010). Pathologic and virologic findings in mid-gestational porcine foetuses
460
after experimental inoculation with PCV2a or PCV2b. Veterinary microbiology 145, 62-68.
461
Segalés, J. (2012). Porcine circovirus type 2 (PCV2) infections: clinical signs, pathology and laboratory
462
463
464
465
466
diagnosis. Virus Res 164, 10-19.
Steinfeldt, T., Finsterbusch, T. & Mankertz, A. (2001). Rep and Rep' protein of porcine circovirus type 1
bind to the origin of replication in vitro. Virology 291, 152-160.
Tischer, I., Gelderblom, H., Vettermann, W. & Koch, M. A. (1982). A very small porcine virus with
circular single-stranded DNA. Nature 295, 64-66.
467
Xiao, C. T., Halbur, P. G. & Opriessnig, T. (2015). Global molecular genetic analysis of porcine circovirus
468
type 2 (PCV2) sequences confirms the presence of four main PCV2 genotypes and reveals a rapid
469
increase of PCV2d. Journal of general virology 96, 1830-1841.
470
Yu, S., Vincent, A., Opriessnig, T., Carpenter, S., Kitikoon, P., Halbur, P. G. & Thacker, E. (2007).
471
Quantification of PCV2 capsid transcript in peripheral blood mononuclear cells (PBMCs) in vitro.
472
Veterinary microbiology 123, 34-42.
473
474
Zinchuk, V. & Zinchuk, O. (2008). Quantitative colocalization analysis of confocal fluorescence microscopy
images. Curr Protoc Cell Biol Chapter 4, Unit 4 19.
475
476
477
478
479
480
481
482
483
484
485
486
487
488
25
489
Figures
490
Fig. 1 (a) Total number of PCV2-positive cells at different time points after transfection.
491
Positive cells were counted in 25 randomly selected fields at 40X. (b) Immunofluorescence
492
stainings of mutants and parental clone by anti-Cap mAb 38C1 at 36 hpt. The data
493
represents three separate experiments performed in duplicate.
494
Fig. 2 (a) Western blot of PCV2 Cap in transfected PK-15 cell lysates at 36 hpt. (b)
495
Kinetics of viral egress as detected by Capture-ELISA of PCV2 capsids in the supernatant
496
of transfected PK-15 cells.
497
Fig. 3 (a) PK-15 cells transfected with mutants or the parental clone, immunostained with
498
anti-Cap mAb 12E12 (red) and anti-Rep mAb F210 D6 (green) at 24 hpt. The images are
499
representative of three separate experiments performed in duplicate. (b) Colocalization
500
quantification of an average of 60 cells per sample studied from three separate experiments
501
performed in duplicate.
502
Fig. 4 (a) Viral titers of the supernatants collected at different time points after transfection.
503
(b) Viral titers of the supernatants collected at different passages.
504
Fig. 5 (a) Cap mRNA copy number kinetics. Figure represents three independent single
505
experiments. (b) Immunostaining of Cap-positive parental clone and M4-transfected
506
PK15-cells. (c) Viral titers of the synonymous M4 mutant and the parental virus in the
26
507
supernatants collected at different time points after transfection. (d) Viral titers of the
508
synonymous M4 mutant and parental viruses at different passages.
509
Fig. 6 (left column) Single transfection with PCV2b-based mutants showing the typical
510
mutant Cap expression pattern. (middle column) Cotransfection of the PCV2b-based
511
mutants with a PCV2a virus strain rescues mutant Cap expression. (right column)
512
Co-transfection does not affect PCV2a-Cap expression.
513
514
27