Requirement of FraG for channel formation between vegetative cells

advertisement
1
Requirement of Fra proteins for communication channels between cells in the filamentous
2
nitrogen-fixing cyanobacterium Anabaena sp. PCC 7120
3
4
5
6
Amin Omairi-Nassera, Vicente Mariscalb, Jotham Austin IIa,c, and Robert Haselkorna*
7
8
a
9
University of Chicago, Chicago, Illinois 60637, USA.
Department of Molecular Genetics and Cell Biology and cAdvanced Electron Microscopy Facility, The
10
b
11
Universidad de Sevilla, E-41092 Seville, Spain.
Instituto de Bioquímica Vegetal y Fotosíntesis, Consejo Superior de Investigaciones Científicas and
12
13
14
A.O.-N. and V.M. contributed equally to this work.
15
*Corresponding author, rh01@uchicago.edu
16
17
18
19
20
21
Author contributions: R.H. designed research and supervised the work; A.O.-N. and V.M. performed research and
analyzed the results; J.A. supervised the work and analyzed data; all authors contributed to writing and editing the paper.
22
1
23
Abstract
24
The filamentous nitrogen-fixing cyanobacterium Anabaena sp. PCC 7120 differentiates
25
specialized cells, heterocysts, that fix atmospheric nitrogen and transfer the fixed nitrogen to
26
adjacent vegetative cells. Reciprocally, vegetative cells transfer fixed carbon to heterocysts.
27
Several routes have been described for metabolite exchange within the filament, one of which
28
involves communicating channels that penetrate the septum between adjacent cells. Several fra
29
gene mutants were isolated 25 years ago on the basis of their phenotypes: inability to fix nitrogen
30
and fragmentation of filaments upon transfer from N+ to N- media. Cryopreservation combined
31
with electron tomography, were used to investigate the role of three fra gene products in channel
32
formation. FraC and FraG are clearly involved in channel formation while FraD has a minor part.
33
Additionally, FraG was located close to the cytoplasmic membrane and in the heterocyst neck,
34
using immunogold labeling with antibody raised to the N-terminal domain of the FraG protein.
35
Significance
36
Cellular communication along the filaments of heterocyst-forming, nitrogen-fixing cyanobacteria
37
has been discussed for at least 50 years but how this might be accomplished is not fully
38
understood. We recently showed that the septum between heterocysts and vegetative cells is
39
pierced by channels 12 nm in diameter and 20 nm long. Here, we show that three proteins, FraC,
40
FraD and FraG, participate in the formation of the channels although none of them appears to be
41
a structural component of the channels. Moreover, using gold particle-labeled antibody, FraG was
42
found around the cyanophycin plug as well as associated with the cytoplasmic membrane in the
43
neighborhood of the peptidoglycan that forms the septum.
2
44
Introduction
45
Cyanobacteria are phototrophic microbes that bear a gram-negative cell envelope and are
46
capable of oxygenic photosynthesis. Some cyanobacteria, such as the filamentous Anabaena sp.
47
strain PCC 7120 (hereafter called Anabaena), are capable of fixing atmospheric N2 when grown
48
in media lacking combined nitrogen. Nitrogen fixation occurs in heterocysts, specialized cells
49
that differentiate from vegetative cells along the filaments and provide a micro-oxic environment
50
for the process (1). One long-standing attraction of Anabaena is its beautiful pattern of
51
differentiation: new heterocysts differentiate midway between two heterocysts as the distance
52
between them doubles due to division of the vegetative cells. This organism, which belongs to
53
one of the first prokaryotic groups on earth to have evolved multicellularity, had to develop
54
structures for intercellular communication. Intercellular communication between heterocysts and
55
vegetative cells comprises small molecules, such as sucrose moving from vegetative cells to
56
heterocysts (2–5) and a dipeptide, -aspartyl-arginine, moving from heterocysts to vegetative
57
cells (6, 7). The mechanism of communication between heterocysts and vegetative cells has been
58
debated for the last 50 years. Two pathways have been proposed for such exchanges (1, 8–10).
59
One is through the periplasm, suggested by the continuity of the outer membrane surrounding the
60
entire filament (9, 11, 12). The other proposed means of communication requires structures
61
between adjacent cells in the filament. Several structures connecting vegetative cells and
62
heterocysts and vegetative cells with each other have been observed using freeze-fracture,
63
conventional electron microscopy and cryo fixation with electron tomography (13–17). Different
64
names have been given to these structures: microplasmodesmata, septosomes, septal junctions, or
65
nanopores (12, 13, 18, 19). Using cryopreservation combined with electron tomography, we
66
observed structures we call channels traversing the peptidoglycan layer in Anabaena (19). These
67
channels are 12 nm long with a diameter of 12 nm, in the septa between vegetative cells. Longer
68
channels, 21 nm long with a similar diameter of 12 nm, were seen in the septa between vegetative
69
cells and heterocysts (20).
70
Several Anabaena gene products were proposed to be involved specifically in intercellular
71
communication. Three were characterized initially from a large set of mutants selected on the
72
basis of their inability to fix nitrogen (21). These mutants manifest a fragmentation phenotype,
73
meaning that they fragment into short filaments upon transfer to liquid medium lacking combined
74
nitrogen, after which they die (15, 22, 23). Further characterization of these mutants led to
3
75
uncovering a role for several fra gene products in intercellular molecular transfer (23–25).
76
fraC encodes a 179-amino acid protein with three predicted trans-membrane segments;
77
fraD encodes a 343-amino acid protein with five predicted trans-membrane segments and a
78
coiled-coil domain; and fraG (also called sepJ) encodes a 751-amino acid protein predicted to
79
have an N-terminal coiled-coil domain, an internal linker domain, and a C-terminal permease-like
80
domain with either 10 trans-membrane segments (22) or 9 or 11 trans-membrane segments (26).
81
fraG deletion prevents heterocyst differentiation and glycolipid layer formation, while the
82
deletion of either fraC or fraD allows heterocyst differentiation, but the heterocysts formed show
83
an aberrant neck and do not fix nitrogen (23, 25). Using GFP tags, FraC, FraD and FraG proteins
84
were shown to be located in the septum between cells (23, 26). FraD was further localized to the
85
septum by immunogold labeling using an antibody raised against the N-terminal coiled-coil part
86
of FraD (25). Fluorescence recovery after photobleaching (FRAP) experiments showed
87
impairment in cell-cell transfer of small molecules such as calcein (622-Da) and 5-
88
carboxyfluorescein (374-Da) in fraC, fraD and fraG mutants, further indicating a role of these
89
gene products in intercellular communication (23–25).
90
In the work reported here, cryopreservation combined with electron tomography was used
91
to investigate the role of these three fra gene products in channel formation. We found that FraC
92
and FraG are clearly required for channel formation, whereas FraD plays a minor role.
93
Immunogold labeling with antibody to the N-terminal coiled-coil domain of FraG yielded an
94
improved localization for FraG.
95
4
96
Results
97
Roles of FraC and FraD in channel formation between vegetative cells
98
In earlier studies, three deletion mutant strains CSVT1 (∆fraC) CSVT2 (∆fraD), and
99
CSVT22 (double mutant ∆fraC/D) revealed the same fragmentation phenotype: upon transfer to
100
N- medium, the filaments fell apart (15, 23, 25). However, in nitrogen-rich media, these mutants
101
did not show any morphological alteration when examined by light microscopy, although transfer
102
of calcein and 5-CFDA between cells was hampered. In addition, GFP fusions to wild-type FraC
103
and FraD allowed localization of both proteins to the septum connecting vegetative cells.
104
Additionally, immunogold labeling of FraD confirmed its location in the septum (23, 25).
105
These observations prompted us to investigate the channels in these mutants under
106
nitrogen-replete conditions. We examined 2.2-nm tomographic sections of septa between
107
vegetative cells of the ∆fraC, ∆fraD and ∆fraC/D mutants in three or four tomographic volumes
108
for each strain, including some tomograms covering the whole septum (serial tomograms). We
109
present here only the middle part of the septum between two vegetative cells (Fig. 1). (For the
110
entire septum see supplement Fig. S1). The septum of CSVT1 (∆fraC) shows fewer and wider
111
channels than the WT (Fig. 1A to 1D). Note that the septum in each tomogram corresponds to a
112
200- to 300-nm section of the entire septum. The dimensions and frequency of channels in the
113
septa of CSVT2 (∆fraD) were similar to those observed in the WT (Fig. 1E and 1F). The septum
114
between vegetative cells in CSVT22 (∆fraC/D) displays only a single channel, this one appearing
115
to be wider than those observed in WT (Fig. 1G and 1H) (other tomograms of CSVT22 show two
116
to three channels in the septum). When the septum is rotated 90° around the y-axis, it is clear that
117
CSVT1 and the double mutant contain about 90% fewer channels than the WT. In ∆fraC, as well
118
as in the double mutant, the length of the channels is 12 nm, which is similar to WT, while the
119
diameter is 21 nm, noticeably wider than WT (Table 1). Based on these observations we conclude
120
that FraC plays a role, possibly structural, in assembly of the channels, while FraD does not.
121
The heterocyst-vegetative cell septa in the ∆fraC/D double mutant
122
Although CSVT22 (∆fraC/D) produces heterocysts in response to nitrogen limitation, it is
123
not able to grow diazotrophically. This phenotype may result from an altered structure of the
124
heterocyst/vegetative cell septa as in the single mutants (25). Four tomograms for the mutant
125
vegetative cell-heterocyst junctions, including one tomogram covering the whole junction (three
5
126
serial tomograms), were analyzed. The cup-like structure typical of the WT heterocyst neck is
127
missing in the CSVT22 strain, consistent with previously reported results (25). In addition, the
128
septum in the ∆fraC/D mutant appears to be three to four times wider than in the WT (82 ±25 nm
129
in the mutant compared to 21 nm in WT; Fig. 2, Table 1). The septum also contains fewer
130
channels than WT, one to four channels per septum, compared to ~20 in WT. These results
131
resemble those for the channels in the septa between vegetative cells of the same mutant, as
132
described in the previous section.
133
134
135
Requirement of FraG for channel formation between vegetative cells
136
tomography. We examined 2.2-nm tomographic sections of septa between vegetative cells of the
137
∆fraG mutant, grown in complete medium, in eight tomographic reconstructions. Only three of
138
the eight reconstructed tomograms showed a few (3-4) channels in the septum (Fig. 3C),
139
compared to 15-20 channels in each of the five WT tomograms. The channels are seen clearly
140
when the tomographic volume is rotated 90° around the y axis, where they appear as white holes
141
in the dark background of the septum. CSVM34 (∆fraG) shows many fewer channels compared
142
to WT (Fig. 3B and 3D). The dimensions of the channels observed in ∆fraG are not statistically
143
different from those of WT (Table 1). These results suggest that FraG might provide a dock for
144
initiating channel assembly, to be discussed below.
We also studied the septum structure in strain CSVM34 (∆fraG) (27), by electron
145
146
FraG is localized around the cyanophycin mass in the heterocyst
147
Previous studies, using GFP-tagged FraG, localized FraG to the intercellular septa, but
148
optical microscopy lacks the resolution needed to define this location precisely. To locate FraG
149
better, we immunolabeled samples of the WT strain (grown with or without combined nitrogen)
150
using antibody raised against the FraG N-terminal coiled-coil domain (anti-FraG_CC) and a 10-
151
nm gold-labeled secondary antibody (27). Very few gold particles were observed in cells grown
152
under N+ conditions (supplement Fig. S2), even in septa where FraG-GFP had been localized by
153
confocal microscopy (26). The specificity of the gold particles in the septum could not be
154
confirmed in comparison with gold particles detected in the background, probably due to the low
155
expression level of FraG. This point will be elaborated in the next paragraph. However, with
156
samples grown under nitrogen-fixing conditions, labeling was observed in the heterocyst neck,
6
157
around the cyanophycin mass, that is, close to the heterocyst-vegetative cell junction. Only a few
158
gold particles were seen on the larger part of the cyanophycin that is close to the polar thylakoid
159
mass. As was the case with cells grown under N+ conditions, very few gold particles were
160
identified in the septum between vegetative cells and between heterocysts and vegetative cells.
161
Figure 4B shows a control experiment in which WT samples were incubated with the secondary
162
gold-labeled antibody alone, without the FraG primary antibody. No gold particles were detected
163
in this control experiment, which indicates that the signal we detect is due to the FraG antibody.
164
The cells shown in Figures 4A and 4B correspond to samples in which the cyanophycin has
165
dropped out during the preparation of the sample. The absence of cyanophycin must have made
166
the FraG-antigenic domain accessible to the antibody in the heterocyst neck. This hypothesis was
167
confirmed by immunogold labeling of cells with intact cyanophycin, which showed very few
168
bound gold particles (supplement Fig. S3).
169
In light of these results, we reinvestigated the C-terminal GFP-tagged FraG localization in
170
heterocysts, using 3D deconvolution fluorescence microscopy in the strain CSAM137 (26). The
171
fluorescence is spread through the outermost part of the heterocyst neck. These results agree with
172
the immunolocalization of FraG in the heterocyst neck around the cyanophycin mass (Fig. 4A).
173
Figure 4F shows two distinguishable spots at each pole of the cell, suggesting the presence of
174
FraG-GFP towards the vegetative cell, as well as around the cyanophycin and more specifically
175
in the part that is present in the heterocyst neck. To further investigate FraG localization around
176
the cyanophycin mass, we collected tomograms for the immunogold-labeled samples (Fig. 5 A-
177
C). The immunotomogram with anti-FraG_CC shows the gold particles around the cyanophycin.
178
Figures 5A to 5C show gold particles at different depths around the cyanophycin mass, between
179
the heterocyst membrane and the cyanophycin. Figure 5D shows the distribution of gold particles
180
around the cyanophycin mass, confirming the localization of FraG in the heterocyst neck. Note
181
that, in the EM images, the cyanophycin always appears split into two parts. The gap is probably
182
due to the cyanophycin breakage that was observed in all heterocysts in this study and in previous
183
studies (20).
184
185
186
7
187
188
189
FraG immunogold localization in the septum between vegetative cells
190
expression of FraG. In order better to localize FraG between vegetative cells, we constructed a
191
mutant, W30, overexpressing FraG. The mutant did not show any phenotypic difference
192
compared to WT (same growth rate in N+ and same pattern and morphology for differentiating
193
heterocysts). Western blots of W30 extracts show a 7-fold increase in the amount of FraG
194
compared to WT (Supplement Fig. S4). Immunogold labeling of W30 using anti-FraG_CC shows
195
the N-terminal domain mainly on the edge of the septa between vegetative cells (Fig. 6A and 6C)
196
distributed close to the cytoplasmic membrane. More than 100 cells were analyzed and all of
197
them show similar gold distribution. (See supplement Fig. S5 for more cells). The septal
198
localization of the native FraG protein corroborates the data obtained with the FraG-GFP fusion
199
protein (26). The specificity of the signal in W30 was confirmed by immunolabeling the deletion
200
mutant CSVM34, in which no signal was detected (Fig. 6B).
The weak gold signals in the WT vegetative cell septa could be explained by the low
201
202
203
204
FraG topology
205
different cyanobacteria predict 9, 10 or 11 trans-membrane segments (28). Our analyses for FraG
206
topology in Anabaena using interPro, an integrated database of predictive protein signatures (29)
207
and Protter (30) supported a 10-transmembrane span model (supplement Fig. S6). This fact
208
affects localization of the coiled-coil domain, which has been described as essential for the
209
function of the protein (26).
Topology prediction of FraG is not clear, since analysis of homologous sequences from
210
To investigate FraG topology experimentally, we fused GFP to the first 391 amino acids
211
of FraG. This sequence includes the N-terminal domain and most of the linker domain of FraG,
212
but excludes all potential trans-membrane domains (Fig. 7A). The construct, called CCL-GFP
213
(coiled-coil-Linker-GFP), was expressed in WT strain yielding strain WGF. Additionally, to
214
exclude any localization due to interaction with the WT FraG, the reporter construct was
215
expressed in a ∆fraG background producing strain ∆GF. Vegetative cells of both WGF and ∆GF
216
show a signal in the division plane, forming a ring similar to the FtsZ division ring (Fig. 7B, 7D
217
and 7E). No signal was detected in the mature septa. Surprisingly, in heterocysts of WGF, the
8
218
CCL-GFP construct was also localized in the poles even though it lacks the predicted trans-
219
membrane domains of FraG (Fig. 7C).
9
220
221
Discussion
222
Septa of several fragmentation mutants that are impaired in intercellular communication
223
were examined by electron tomography. None of the fraC, fraD and fraG mutants showed a total
224
loss of channels in their septa, but rather a decrease in the number of channels and different sizes
225
for these channels were observed, especially for ∆fraC and ∆fraG, indicating that FraC and FraG
226
either affect the assembly of the channels or are involved in the assembly of different channels.
227
The difference in channel distribution between CSVT1 and CSVT2 suggest different
228
functions for FraC and FraD, respectively. fraC deletion results in the reduction of channel
229
number in the septum, which suggests that FraC is either a part of the channel structure or a
230
regulator of channel assembly. On the other hand, fraD deletion does not seem to affect channel
231
formation between vegetative cells. FraD is probably involved in maintaining a stable cell-cell
232
contact in the septum. We cannot exclude a role for FraD in recruiting components of other types
233
of channels that are not detected using our methods. The ∆fraC/D double mutant (CSVT22)
234
showed a 7-fold increase of the width of the septa between vegetative cells and heterocysts
235
compared to the 2-fold increase observed in corresponding WT septa. These results suggest that
236
FraC or FraD or both play a role in expansion of the peptidoglycan and the channels that connect
237
heterocysts with vegetative cells. Aberrant heterocyst neck structures were previously shown in
238
fraC and fraD mutants (25), suggesting that FraC and FraD are important for maintaining a tight
239
junction at the septum during the restructuring or remodeling of the peptidoglycan layer
240
throughout the heterocyst differentiation process. The heterocyst neck in these mutants lacks the
241
typical cup-like structure found in the WT. The increase in the septum width and the decrease in
242
the contact area between the heterocyst and the vegetative cell could explain the fragmentation
243
phenotype of the fraC, fraD and fraC/D mutants (15, 23, 25).
244
245
Tomograms for the septa between vegetative cells of CSVM34 (∆fraG), using cells fixed
246
with potassium permanganate, were previously analyzed (12). Structures called “septosomes”,
247
were measured to be 18 nm long in CSVM34 compared to 27 nm in WT. The septosome
248
frequency was difficult to measure due to lack of resolution. Using cryo-fixation and staining
249
with osmium tetroxide, which highlights peptidoglycan, and two-axis tomograms, we observed
10
250
channels with similar length in WT and CSVM34 although there were fewer channels in
251
CSVM34. This difference in observations suggests that different structures connecting cells
252
might be revealed when using cryo-fixation and tomography compared to the structures observed
253
in cells fixed with potassium permanganate (12). However, in agreement with our observations,
254
strain CSVM34 shows a reduced number of septal peptidoglycan nanopores (15% of the wild
255
type) (31).
256
FraG-GFP has been seen near the septum between vegetative cells (Flores et al., 2007). In
257
this work we have improved the resolution of the localization of FraG by means of immunogold
258
labeling. The N-terminal coiled-coil domain of FraG was detected close to the septum between
259
vegetative cells and in the heterocyst, in the neck around the cyanophycin mass. The positions of
260
the gold particles around the cyanophycin mass in heterocysts assign the location of the N-
261
terminal domain of FraG to a position facing the cyanophycin cavity. We cannot exclude the
262
possibility that FraG is anchored to a putative hydrophobic layer surrounding the cyanophycin
263
plug. However, due to the fact that the gold particles could be ~30 nm from their antigenic target
264
(32), the trans-membrane domains of FraG could be located in the cytoplasmic membrane with
265
its N-terminal coiled-coil domain facing toward the cyanophycin plug, as modeled in Figure 8.
266
In heterocysts, the (artefactual) lost cyanophycin apparently made FraG antigens
267
accessible to gold labeling; FraG antibodies reacted with their respective antigen within the 200-
268
nm section. In the case of vegetative cells and that of heterocysts with intact cyanophycin,
269
immunogold-labeled antibodies react only with their respective antigens that are located on the
270
surface of each section. Some gold particles can be located on the surface near the septum
271
between vegetative cells (Supplement Fig. S2) and between heterocysts and vegetative cells (Fig.
272
4A and supplement Fig. S3). In these areas, the gold-labeled secondary antibody is probably
273
bound to exposed FraG coiled coils. The distribution of the gold particles around the cytoplasmic
274
membrane between vegetative cells (Fig. 6A and 6C) suggests a localization of the N-terminal
275
closer to the cytoplasmic membrane of the cells rather than the peptidoglycan.
276
It has been established in bacteria that GFP fused to the C-terminal domain of cytoplasmic
277
membrane proteins fluoresce when the GFP is located in the cytoplasm (33), or in the periplasm
278
when it is exported by the TAT system (9). The C-terminal GFP-tagged FraG shows its
279
fluorescence signal around the heterocyst neck (Fig. 4), which indicates that the FraG C-terminal
280
domain is located in the cytoplasm. Moreover, the C-terminal GFP-tagged FraG (CCL-GFP)
11
281
containing the coiled-coil and part of the linker domain of FraG and lacking the trans-membrane
282
domain (Fig. 7A), shows its fluorescence signal in a ring in dividing vegetative cells and in the
283
heterocyst poles (Fig. 7). Since a TAT signal peptide is not detected in the N-terminal sequence
284
of FraG, GFP fluorescence indicates that the FraG N-terminal domain might be located in the
285
cytoplasm. The absence of GFP signals in all septa between vegetative cells in WGF and GF is
286
probably due to the absence of the trans-membrane domain that anchors the protein to the plasma
287
membrane. The GFP fluorescence signal in the division planes of vegetative cells in the CCL-
288
GFP construct are in agreement with recent results showing that FraG interacts with FtsQ, a
289
protein involved in cell division (34). On the other hand, the presence of the GFP signal in WGF
290
within the cyanophycin plug could be due to interaction of the CCL-GFP construct with either
291
WT FraG or the cyanophycin.
292
The Anabaena open reading frame alr2338 was first denoted fraG (22). Later, sepJ was
293
preferred because it better reflected the subcellular localization found in (26). In this work, we
294
show that the alr2338 product is not located exclusively in the septum. Therefore, we prefer fraG
295
for that gene, reflecting its phenotype.
296
12
297
Experimental procedures
298
299
Anabaena strains and growth conditions
300
Anabaena sp. strain PCC 7120 and derivative strains were grown photoautotrophically at
301
30 °C under constant white light (35 μE m-2 s-1), in a CO2 enriched atmosphere. The medium
302
composition is similar to BG11 with some modification (35). For heterocyst induction, cells were
303
harvested by centrifugation and resuspended in NO3- deficient medium, for which NaNO3 was
304
replaced by NaCl. Mutant strains CSVT1 (∆fraC), CSVT2 (∆fraD), CSVT22 (∆fraC/D), and
305
CSVM34 (∆fraG) have been described previously (23, 25, 27).
306
supplemented when appropriate with Neomycin (50 g mL-1). Strain W30 was constructed by
307
transferring pAN130 (22), a replicative plasmid containing the promoter and coding region of
308
fraG to WT Anabaena by conjugation as described (36).
Growth media were
309
In order to express the CC-linker GFP-tagged FraG, a 1941-bp fragment covering the 5’
310
non-coding region and part of the fraG orf was amplified using primers FFB
311
(GGATCCTGAAATATGAGTTATGGCTGGGGAC)
312
GGCGGAGGAGTTG), which also creates BamHI and NheI restriction sites, respectively. DNA
313
from Anabaena sp. PCC7120 was used as template. The PCR product, digested with BamHI and
314
NheI, was cloned into pRL25N (37) digested with the same enzymes, yielding pRCG. The
315
plasmid was sequenced to verify the fidelity of the PCR. pRCG was transferred to WT and to
316
CSVM34 by conjugation as described above, yielding WGF and ∆GF, respectively.
317
Embedding Anabaena for electron microscopy
and
FRN
(GCTaGcTGGTGCA
318
Anabaena cells were harvested by centrifugation and transferred to an aluminum sample
319
holder and cryoprotected with 0.15 M sucrose. Samples were frozen in a Baltec HPM 010 high-
320
pressure freezer, then freeze-substituted in 2% osmium tetroxide (EMS), using an Automated
321
Freezing Substitution machine (ASF2, Leica), in anhydrous acetone at -80 C for 72 h. The
322
temperature was then increased from -80 C to -20 C over 12 h. Samples were then washed with
323
acetone three times at -20 C, then transferred to 4 C, held overnight, and then warmed to room
324
temperature. Samples were then infiltrated with increasing concentrations of EPON resin (5%,
13
325
10%, 25%, 50%, 75% and 100%) finally polymerized at 60 C for 24 h (see (38) for more
326
details).
327
For immunogold labeling, the high-pressure frozen samples were substituted in 0.1%
328
uranyl acetate in acetone at -80°C for 3 days and then warmed to -50°C for 12 h. After three
329
acetone rinses, samples were slowly infiltrated under controlled time and temperature conditions
330
in a Leica AFS system at -50°C with Lowicryl HM20 resin according to the following schedule:
331
5, 10, 25, 50, 75, and 100% (24 h incubation for each concentration). After the last incubation
332
with 100% HM20, samples were rinsed with fresh 100% HM20 three times, with 1h for each
333
wash. Samples were finally polymerized at the same temperature under UV light for 32 h.
334
Immunocytochemistry
335
336
Rabbit antibodies raised against the 188-amino acid coiled-coil domain of FraG were used
337
to detect FraG in Anabaena heterocysts and vegetative cells. Samples embedded in Lowicryl
338
HM20 were cut into 150-nm-thick sections and placed on Formvar-coated gold slot grids.
339
Immunocytochemistry was performed essentially as described by Otegui et al. (39). Sections
340
were blocked for 20 min with a 5% (w/v) solution of nonfat milk in TBS plus 0.1%Tween 20
341
(TBST). Primary antibodies were diluted 1:20 in a solution of 2.5% nonfat milk in TBST at room
342
temperature for 1 h. The sections were rinsed in a stream of TBS plus 0.5% Tween 20 and then
343
transferred to the secondary antibody (goat anti-rabbit IgG 1:20 in TBST) conjugated to 10-nm
344
gold particles, for 1 h. Control procedures omitted the primary antibody. For gold quantification,
345
over 1000 gold particles were counted in more than 30 pairs of cells. See supplement Fig. S7, for
346
the calculation of gold distribution in the septum vs the rest of the cell and the background.
347
Sectioning
348
349
EPON sections (100 to 300 nm) were cut using a Leica EM AFS2 Automatic Freeze-
350
Substitution Processor and collected on 1% Formvar (EMS) copper slot grids. Sections were
351
stained with 2% uranyl acetate and 0.5% lead citrate for 8 and 5 min, respectively. For tomogram
352
collection, 300-nm sections were used and 10 L of 15-nm colloidal gold (BBI solutions) were
353
applied for 10 min on each side of the grid as fiduciary markers.
14
354
Electron Tomography
355
356
Tomograms were collected using a Tecnai G2 F30 (FEI) electron microscope operating at
357
300 kV. Images were taken at 15,000x from -60 to +60 with 1 interval. Each tomogram was
358
collected in two perpendicular axes. Etomo was used to build the tomograms and to merge the
359
two single-axis tomograms into one dual-axis tomogram. Tomograms were then displaced,
360
analyzed and modeled using the 3DMOD software (40).
361
Fluorescence microscopy
362
363
Anabaena cells were visualized with a Leica DM6000B fluorescence microscope and an
364
ORCA-ER camera (Hamamatsu) using an FITC L5 filter (excitation, band-pass [BP] 480/40
365
filter; emission, BP 527/30 filter) and the Leica SP5 2-photon confocal microscope. The images,
366
including BlindDeblur deconvolution of 3D images, were produced using the LAS AF Leica
367
software.
368
369
Preparation of Cell Extracts and Western blots.
370
Total cell extracts were prepared as described in (41). Proteins were separated using
371
Novex 14% Tris-Glycine gels (Novex, Life Technology). Chlorophyll concentration was used to
372
ensure equivalent loading of cell extracts. 1.2 g chl was loaded per 1-mm well for blotting and
373
Coomassie staining.
374
For immunoblots, proteins were transferred to PVDF membranes (immobilon, Millipore)
375
using a Bio-Rad gel transfer system (Bio-Rad). Blots were blocked with Tris-buffered saline
376
supplemented with 0.1% Tween and 5% dry skimmed milk and incubated with the primary
377
antibody (1:500 dilution for FraG and 1:10,000 dilution for FNR) over-night at 4 °C. After
378
washing, blots were incubated 1 h at room temperature with a 1:15,000 dilution of peroxidase-
379
conjugated anti-rabbit IgG (Promega, Madison, WI). The signal was visualized using ECL
380
chemiluminescent substrate (SuperSignal West Pico Chemiluminescent, Thermo Scientific).
381
Images were generated with a CCD camera and analyzed using ImageJ software.
382
383
15
384
Acknowledgments. This work was supported by the Ellison Medical Foundation and The
385
University of Chicago. We thank Prof. Enrique Flores for support in part from grant no.
386
BFU2011-22762 from Plan Nacional de Investigación, Spain, co-financed by the European
387
Regional Development Fund. We also thank Sean Callahan for plasmid pAN130 and Amel Latifi
388
for plasmid pRL25N. William Buikema and Ghada Ajlani provided critical reading of the paper.
389
390
16
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
Haselkorn R (2008) Cell–cell communication in filamentous cyanobacteria. Mol Microbiol
70(4):783–785.
Jüttner F (1983) 14C-labeled metabolites in heterocysts and vegetative cells of Anabaena
cylindrica filaments and their presumptive function as transport vehicles of organic carbon
and nitrogen. J Bacteriol 155(2):628–33.
Cumino AC, Marcozzi C, Barreiro R, Salerno GL (2007) Carbon Cycling in Anabaena sp.
PCC 7120. Sucrose Synthesis in the Heterocysts and Possible Role in Nitrogen Fixation.
Plant Physiol 143(3):1385–1397.
López-Igual R, Flores E, Herrero A (2010) Inactivation of a Heterocyst-Specific Invertase
Indicates a Principal Role of Sucrose Catabolism in Heterocysts of Anabaena sp. J Bacteriol
192(20):5526–5533.
Vargas WA, Nishi CN, Giarrocco LE, Salerno GL (2010) Differential roles of
alkaline/neutral invertases in Nostoc sp. PCC 7120: Inv-B isoform is essential for
diazotrophic growth. Planta 233(1):153–162.
Hejazi M, et al. (2002) Isoaspartyl dipeptidase activity of plant-type asparaginases. Biochem
J 364(Pt 1):129–136.
Burnat M, Herrero A, Flores E (2014) Compartmentalized cyanophycin metabolism in the
diazotrophic filaments of a heterocyst-forming cyanobacterium. Proc Natl Acad
Sci:201318564.
Flores E, Herrero A (2010) Compartmentalized function through cell differentiation in
filamentous cyanobacteria. Nat Rev Microbiol 8(1):39–50.
Mariscal V, Herrero A, Flores E (2007) Continuous periplasm in a filamentous, heterocystforming cyanobacterium. Mol Microbiol 65(4):1139–1145.
Mariscal V, Flores E (2010) Multicellularity in a heterocyst-forming cyanobacterium:
pathways for intercellular communication. Adv Exp Med Biol 675:123–135.
Flores E, Herrero A, Wolk CP, Maldener I (2006) Is the periplasm continuous in
filamentous multicellular cyanobacteria? Trends Microbiol 14(10):439–443.
Wilk L, et al. (2011) Outer membrane continuity and septosome formation between
vegetative cells in the filaments of Anabaena sp. PCC 7120. Cell Microbiol 13(11):1744–
1754.
Giddings TH, Staehelin LA (1978) Plasma membrane architecture of Anabaena cylindrica:
occurrence of microplasmodesmata and changes associated with heterocyst development
and the cell cycle. Eur J Cell Biol (16):235–249.
Giddings TH, Staehelin LA (1981) Observation of microplasmodesmata in both heterocystforming and non-heterocyst forming filamentous cyanobacteria by freeze-fracture electron
microscopy. Arch Microbiol 129(4):295–298.
Bauer CC, Buikema WJ, Black K, Haselkorn R (1995) A short-filament mutant of
Anabaena sp. strain PCC 7120 that fragments in nitrogen-deficient medium. J Bacteriol
177(6):1520–1526.
Lang NJ, Fay P (1971) The Heterocysts of Blue-Green Algae. II. Details of Ultrastructure.
Proc R Soc Lond B Biol Sci 178(1051):193–203.
Wildon D, Mercer F (1963) The Ultrastructure of the Vegetative Cell of Blue-Green Algae.
Aust J Biol Sci 16(3):585–596.
17
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
18. Lehner J, et al. (2013) Prokaryotic multicellularity: a nanopore array for bacterial cell
communication. FASEB J Off Publ Fed Am Soc Exp Biol 27(6):2293–2300.
19. Mariscal V (2014) Cell-Cell joining proteins in heterocyst-forming cyanobacteria. In:
Flores, E. and Herrero, A. (eds) The Cell Biology of Cyanobacteria. Caister Academic
Press. pp. 293-304.
20. Omairi-Nasser A, Haselkorn R, Austin J (2014) Visualization of channels connecting cells
in filamentous nitrogen-fixing cyanobacteria. FASEB J:fj.14–252007.
21. Buikema WJ, Haselkorn R (1991) Isolation and complementation of nitrogen fixation
mutants of the cyanobacterium Anabaena sp. strain PCC 7120. J Bacteriol 173(6):1879–
1885.
22. Nayar AS, Yamaura H, Rajagopalan R, Risser DD, Callahan SM (2007) FraG is necessary
for filament integrity and heterocyst maturation in the cyanobacterium Anabaena sp. strain
PCC 7120. Microbiology 153(2):601–607.
23. Merino-Puerto V, Mariscal V, Mullineaux CW, Herrero A, Flores E (2010) Fra proteins
influencing filament integrity, diazotrophy and localization of septal protein SepJ in the
heterocyst-forming cyanobacterium Anabaena sp. Mol Microbiol 75(5):1159–1170.
24. Mullineaux CW, et al. (2008) Mechanism of intercellular molecular exchange in heterocystforming cyanobacteria. EMBO J 27(9):1299–1308.
25. Merino-Puerto V, et al. (2011) FraC/FraD-dependent intercellular molecular exchange in the
filaments of a heterocyst-forming cyanobacterium, Anabaena sp. Mol Microbiol 82(1):87–
98.
26. Flores E, et al. (2007) Septum-Localized Protein Required for Filament Integrity and
Diazotrophy in the Heterocyst-Forming Cyanobacterium Anabaena sp. Strain PCC 7120. J
Bacteriol 189(10):3884–3890.
27. Mariscal V, Herrero A, Nenninger A, Mullineaux CW, Flores E (2011) Functional
dissection of the three-domain SepJ protein joining the cells in cyanobacterial trichomes.
Mol Microbiol 79(4):1077–1088.
28. Nürnberg DJ, et al. (2014) Branching and intercellular communication in the Section V
cyanobacterium Mastigocladus laminosus, a complex multicellular prokaryote. Mol
Microbiol 91(5):935–949.
29. Mitchell A, et al. (2015) The InterPro protein families database: the classification resource
after 15 years. Nucleic Acids Res 43(Database issue):D213–221.
30. Omasits U, Ahrens CH, Müller S, Wollscheid B (2013) Protter: interactive protein feature
visualization and integration with experimental proteomic data. Bioinformatics:btt607.
31. Nürnberg DJ, et al. (2015) Intercellular Diffusion of a Fluorescent Sucrose Analog via the
Septal Junctions in a Filamentous Cyanobacterium. mBio 6(2):e02109–14.
32. Hermann R, Walther P, Müller M (1996) Immunogold labeling in scanning electron
microscopy. Histochem Cell Biol 106(1):31–39.
33. Drew D, et al. (2002) Rapid topology mapping of Escherichia coli inner-membrane proteins
by prediction and PhoA/GFP fusion analysis. Proc Natl Acad Sci 99(5):2690–2695.
34. Ramos-León F, Mariscal V, Frías JE, Flores E, Herrero A (2015) Divisome-dependent
subcellular localization of cell-cell joining protein SepJ in the filamentous cyanobacterium
Anabaena. Mol Microbiol:n/a–n/a.
35. Ughy B, Ajlani G (2004) Phycobilisome rod mutants in Synechocystis sp. strain PCC6803.
Microbiology 150(Pt 12):4147–56.
18
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
36. Elhai J, Vepritskiy A, Muro-Pastor AM, Flores E, Wolk CP (1997) Reduction of conjugal
transfer efficiency by three restriction activities of Anabaena sp. strain PCC 7120. J
Bacteriol 179(6):1998–2005.
37. Zhang L-C, Risoul V, Latifi A, Christie JM, Zhang C-C (2013) Exploring the size limit of
protein diffusion through the periplasm in cyanobacterium Anabaena sp. PCC 7120 using
the 13 kDa iLOV fluorescent protein. Res Microbiol 164(7):710–717.
38. Austin JR 2nd (2014) High-Pressure Freezing and Freeze Substitution of Arabidopsis for
Electron Microscopy. Arabidopsis Protocols, Methods in Molecular Biology., eds SanchezSerrano JJ, Salinas J (Humana Press), pp 473–486.
39. Otegui MS, Mastronarde DN, Kang B-H, Bednarek SY, Staehelin LA (2001) ThreeDimensional Analysis of Syncytial-Type Cell Plates during Endosperm Cellularization
Visualized by High Resolution Electron Tomography. Plant Cell 13(9):2033–2052.
40. Kremer JR, Mastronarde DN, McIntosh JR (1996) Computer visualization of threedimensional image data using IMOD. J Struct Biol 116(1):71–76.
41. Omairi-Nasser A, de Gracia AG, Ajlani G (2011) A larger transcript is required for the
synthesis of the smaller isoform of ferredoxin:NADP oxidoreductase. Mol Microbiol
81(5):1178–1189.
19
499
Figures
500
501
Figure 1. The septum between two vegetative cells of WT Anabaena and three
502
fragmentation mutants. A) Electron tomographic image of the WT septum. B) Septum
503
shown in “A” is rotated 90° around the y axis showing the channel distribution within the
504
septum. Several channels are observed in the middle of the septum (white holes). C)
505
Electron tomographic image of the septum of CSVT1 mutant (∆fraC). The septum contains
506
fewer channels compared to WT. D) Septum shown in C is rotated 90° around the y axis. E)
507
Electron tomographic image of the septum of CSVT2 mutant (∆fraD). The septum contains
508
similar number of channels compared to WT. F) Septum shown in E is rotated 90° around
509
the y axis. G) Electron tomographic image of the septum of CSVT22 mutant (∆fraC/D). H)
510
Septum shown in G is rotated 90° around the y axis showing the channel distribution.
511
Arrowheads indicate the channels observed on their corresponding panel before rotation.
512
“t” indicates thylakoids.
513
tomographic slices. Scale bar 50 nm
All images are composed of 10 superimposed 2.2 nm serial
514
515
Figure 2. Heterocyst-vegetative cell septa in WT and fragmentation mutants. A) Electron
516
tomographic image of a WT heterocyst junction. White arrowheads point to the edges of the
517
septum. Yellow arrowheads show the channels that connect the heterocyst and the
518
vegetative cell. B) Electron tomographic image of the CSVT22 (fraCD) heterocyst junction.
519
The septum in this mutant is thicker and only 1-2 channels are present compared to WT.
520
The yellow arrow points to the only channel observed in this tomogram. Black arrowheads
521
point to the plasma membrane in vegetative cells in each panel.
522
All tomographic images are composed of 10 superimposed 2.2-nm tomographic slices. Het:
523
heterocyst. Veg: vegetative cell. Scale bar: 200 nm.
524
525
Figure 3. The septum between two vegetative cells of WT Anabaena and of fragmentation
526
mutants. A) Electron tomographic image of the WT septum. B) Septum shown in “A” is
527
rotated 90° around the y axis showing the channel distribution within the septum. Several
528
channels are observed in the middle of the septum (white holes). C) Electron tomographic
20
529
image for the septum of mutant CSVM34 (∆fraG). D) Septum shown in C is rotated 90°
530
around the y axis showing the channel distribution within the septum; only 2 channels are
531
observed compared to 15 to 20 in WT. Arrowheads in B and D indicate channels observed
532
in the septum in panels A and C, respectively. “t” indicates thylakoids. All images are
533
composed of 10 superimposed 2.2 nm serial tomographic slices. Scale bar 50 nm
534
535
Figure 4. Subcellular localization of FraG in Anabaena heterocysts. A) Immunogold
536
labelling of WT Anabaena using antibodies (black dots) raised against the N-terminal
537
coiled-coil domain of FraG. B) Control immunogold labeling of WT using only secondary
538
antibody; no dots. C) Light transmission micrograph of WT Anabaena grown under N-
539
conditions. D) Autofluorescence of the same cells shown in C. Heterocysts do not show
540
autofluorescence due to loss of PS II chlorophyll. E) Light transmission micrograph of the
541
CSAM137 mutant (FraG-GFP) grown under N- conditions. F) Autofluorescence (red) and
542
GFP fluorescence (green) of the same cells shown in E. GFP fluorescence locates FraG at the
543
poles of the heterocysts. C: Cyanophycin. Het: heterocyst. Veg: vegetative cell. Scale bar: 200
544
nm.
545
546
Figure 5. Serial immunoelectron tomography of WT Anabaena grown under N- conditions.
547
(A) to (C) Serial 2.2-nm tomographic slice images (every 50th slice) through a heterocyst
548
neck labeled with anti-FraG antibody. Note that as one proceeds from the top of the section
549
(A) to the bottom (C), different groups of 10-nm gold labels, indicated by arrows and
550
numbers, are seen at different depths within the cyanophycin interior; 1 and 2 at the top
551
(A), 3 in the middle (B) and 4 at the bottom (C) of the cyanophycin plug. D) Tomographic
552
model of the immunoelectron tomogram showing the location of FraG around the
553
cyanophycin (blue). Green arrows indicate gold particles in the cyanophycin at different
554
depth with n° 1 and 2 seen in section 10 (A), n° 3 in section 60 (B) and n° 4 in section 110
555
(C). Note that there is no contact between the gold particle and the channels (or
556
peptidoglycan). Peptidoglycan layer is red. Het: heterocyst. Veg: vegetative cell. Scale bar:
557
100 nm.
558
21
559
Figure 6. FraG localization between vegetative cells. (A) Immunogold labeling of W30,
560
overexpressing FraG, using anti-FraG antibody. (B) Immunogold labeling of CSVM34
561
(fraG) shows no gold particles. (C) Zoom in to the septum in A; eleven gold particles seen
562
on both sides of the septum. Scale bar: 200 nm.
22
563
Figure 7. FraG N-terminal localization. (A) Cartoon showing the different domains of FraG
564
and the location of the GFP insertion in the linker domain of FraG in the pRGF plasmid. The
565
plasmid was introduced into both WT and ∆fraG yielding WGF and ∆GF. (B) and (C)
566
Autofluorescence (red) and GFP fluorescence (green) in WGF grown in N+ and N-
567
respectively. (D) Autofluorescence (red) and GFP fluorescence (green) in W30 grown in N+
568
(Note that the mutant cannot grow in N- media). (E) Same micrograph shown in D rotated
569
45 around the y axis and showing only GFP fluorescence. Notched arrowheads indicate the
570
location of the CCL-GFP construct in the divisome plane. Straight arrowheads indicate cells
571
at the end of division, hence the presence of a GFP signal. GFP fluorescence shows the FraG
572
N-terminal-linker domain as rings in the divisome plane of vegetative cells. CC; predicted
573
Coiled Coil. L; predicted Linker domain. TM; Trans-membrane domain. N-ter; N-terminal
574
domain. C ter; C terminal domain. M1; FraG 1st Methionine. P391; Proline- the 391st amino
575
acid in FraG linker domain where GFP was fused. Scale bar: 2 m.
576
577
Figure 8. Model for the heterocyst-vegetative cell junction showing putative localization
578
and interactions of FraC, FraD and FraG. These proteins appear to be located in the plasma
579
membrane and/or septum and involved in channel formation, either directly or by
580
recruiting other factors. In a fully-developed heterocyst, FraG is found in the heterocyst
581
neck around the cyanophycin, implicating FraG in an additional role to channel formation,
582
possibly assembly or maintenance of heterocyst neck formation. CC; predicted Coiled Coil.
583
L; predicted Linker domain. MD; Trans-membrane domain. PM; Plasma membrane. PG;
584
Peptidoglycan. OM; Outer membrane. C; Cyanophycin.
585
23
Download