© 2014. Published by The Company of Biologists Ltd.

1

Nephrocystin-4 controls ciliary trafficking of membrane and large soluble proteins at the

2

transition zone

3 4

Junya Awata1, Saeko Takada1,4, Clive Standley3, Karl F. Lechtreck1,5, Karl D. Bellvé3, Gregory

5

J. Pazour2, Kevin E. Fogarty3 and George B. Witman1

Journal of Cell Science

Accepted manuscript

6 7

1

Department of Cell and Developmental Biology, Program in Molecular Medicine, and

8

3

Biomedical Imaging Group, University of Massachusetts Medical School, Worcester MA 01605

9

4

Department of Biochemistry, Molecular Biology, and Biophysics, University of Minnesota,

10 11

2

Minneapolis, MN 55455 5

Department of Cellular Biology, University of Georgia, Athens, GA 30602

12 13 14 15 16

Correspondence to:

17 18

George B. Witman

19

Department of Cell and Developmental Biology

20

University of Massachusetts Medical School

21

55 Lake Avenue North

22

Worcester, MA 01655

23 24

TEL: 508-856-4038

25

FAX: 508-856-1033

26

E-mail: [email protected]

27 28 29

1 JCS Advance Online Article. Posted on 22 August 2014

Accepted manuscript

ABSTRACT

31

The protein nephrocystin-4 (NPHP4) is widespread in ciliated organisms, and defects in NPHP4

32

cause nephronophthisis and blindness in humans. To learn more about NPHP4’s function, we

33

have studied it in Chlamydomonas reinhardtii. NPHP4 is stably incorporated into the distal part

34

of the flagellar transition zone, close to the membrane and distal to CEP290, another transition

35

zone protein. Therefore, these two proteins, which are incorporated into the transition zone

36

independently of each other, define different domains of the transition zone. A nphp4 null

37

mutant forms flagella with nearly normal length, ultrastructure, and intraflagellar transport.

38

When fractions from isolated wild-type and nphp4 flagella were compared, few differences were

39

observed between the axonemes, but a subset of membrane proteins was greatly reduced in the

40

mutant flagella, and cellular housekeeping proteins >50 kDa were no longer excluded from

41

mutant flagella. Therefore, NPHP4 functions at the transition zone as an essential part of a

42

barrier that regulates both membrane and soluble protein composition of flagella. The

43

phenotypic consequences of NPHP4 mutations in humans likely follow from protein

44

mislocalization due to defects in the TZ barrier.

Journal of Cell Science

30

2

45

INTRODUCTION

Journal of Cell Science

Accepted manuscript

46 47

Nephronophthisis (NPHP) is a recessive form of cystic kidney disease that is the most frequent

48

genetic cause of chronic renal failure in children (Wolf and Hildebrandt, 2011). To date, 18

49

causative human genes for NPHP have been reported (Wolf and Hildebrandt, 2011; Failler et al.,

50

2014; Renkema et al., 2014); one of these is NPHP4, which encodes nephrocystin-4 (NPHP4).

51

Mutations in NPHP4 lead to juvenile NPHP type 4 as well as retinitis pigmentosa; both disease

52

features occur together in Senior-Loken syndrome (Hoefele et al., 2005; Mollet et al., 2002; Otto

53

et al., 2002; Schuermann et al., 2002).

54

NPHP and Senior-Loken syndrome are ciliopathies (diseases caused by defects in cilia),

55

and there is abundant evidence from model systems that loss of NPHP4 causes a ciliary

56

phenotype. A study of Nphp4 mutant mice showed that Nphp4 is necessary for normal

57

photoreceptor formation and maintenance as well as sperm development (Won et al., 2011). In

58

zebrafish, nphp4 morphants exhibited classic ciliary phenotypes, including abnormal body

59

curvature and pronephric cysts (Burcklé et al., 2011; Slanchev et al., 2011). In Caenorhabditis

60

elegans, RNAi knockdown of NPHP-4 caused defects in male mating behavior, which is

61

mediated by sensory cilia (Wolf et al., 2005), and a mutant null for NPHP-4 had mild defects in

62

cilia-dependent sensory functions (Winkelbauer et al., 2005). Cilia of the same mutant were

63

stunted, misshapen, and had axonemal ultrastructural defects (Jauregui et al., 2008).

64

These phenotypes most likely result from defective function of NPHP4 at the ciliary

65

transition zone (TZ), a specialized region between the basal body and the cilium proper (Shiba

66

and Yokoyama, 2012; Czarnecki and Shah, 2012; Reiter et al., 2012). Several studies have

67

reported that NPHP4 localizes to the TZ. LAP-tagged NPHP4 in IMCD3 cells is present at sites

68

of cell-cell contact and at the base of the primary cilium, specifically in the TZ (Sang et al.,

69

2011). Others have reported NPHP4 in the TZ of cultured mouse renal epithelial cells (Shiba et

70

al., 2010) and in the connecting cilium -- a greatly elongated TZ -- of mouse photoreceptor cells

71

(Roepman et al., 2005; Patil et al., 2012; Won et al., 2011). In Chlamydomonas reinhardtii,

72

NPHP4 was identified in a proteomic study of basal bodies, which included the TZ (Keller et al.,

73

2005). In C. elegans, NPHP4 has been localized to the TZ of the sensory cilia (Winkelbauer et

74

al., 2005; Jauregui et al., 2008; Williams et al., 2011).

3

Journal of Cell Science

Accepted manuscript

75

Evidence is accumulating that the TZ functions as the “pore” or gate proposed to exist at

76

the base of the cilium to control access to the ciliary compartment (Rosenbaum and Witman,

77

2002). A highly conserved feature of the TZ is the presence of Y-shaped links (Y-links) that

78

connect the TZ doublet microtubules to the overlying membrane (Fisch and Dupuis-Williams,

79

2011). In C. reinhardtii, these links are partially disrupted in the absence of CEP290, another TZ

80

protein, resulting in abnormal flagellar protein composition (Craige et al., 2010). The mutant

81

also has defects in intraflagellar transport (IFT), which is used by nearly all organisms to build

82

cilia and flagella (Pedersen and Rosenbaum, 2008; Ishikawa and Marshall, 2011). These results

83

provided experimental evidence that the TZ functions to control protein and IFT particle entry

84

into the cilium, and indicated that CEP290 is important for this function. NPHP4 also may be

85

involved in this activity: Cilia of the C. elegans nphp-4 mutant accumulate the membrane-

86

associated proteins RGI-2 and TRAM-1a, which normally are excluded from the cilium

87

(Williams et al., 2011).

88

Despite this progress, we still do not know NPHP4’s precise location in the TZ nor how

89

its loss affects overall ciliary composition. In this study, we have used C. reinhardtii to learn

90

more about NPHP4. C. reinhardtii has many advantages for studying cilia and ciliary

91

components. Particularly relevant for this study, its flagella can be isolated to determine the

92

biochemical consequences of loss of NPHP4. NPHP4 is highly conserved (C. reinhardtii to

93

human BLASTN E value = 1e-74), so conclusions from studying C. reinhardtii NPHP4 are

94

likely to be applicable to humans and other organisms. We found that NPHP4 is located at the

95

periphery of the distal TZ, close to the membrane and distal to CEP290. In contrast to CEP290,

96

NPHP4 at the TZ does not undergo rapid turnover. We identified an NPHP4 null mutant; the

97

mutant has full-length flagella and generally has normal TZ and axonemal ultrastructure. We

98

isolated the nphp4 mutant flagella and found that a subset of membrane-associated proteins,

99

present in wild-type flagella, are greatly decreased in amount; conversely, the flagella contain

100

many large cytosolic housekeeping proteins that normally are excluded from wild-type flagella.

101

The results indicate that NPHP4 is a critical component of the selective gate that functions at the

102

TZ to control movement of both soluble and membrane-associated proteins between the flagellar

103

and cytoplasmic compartments. It is likely that the various phenotypic consequences of NPHP4

104

mutations in humans and other organisms all follow from protein mislocalization due to defects

105

in the TZ barrier.

4

106

RESULTS

Accepted manuscript

107 108

NPHP4 loss has minor effects on cell motility but slows flagellar assembly

109

To identify a C. reinhardtii nphp4 mutant, genomic DNA from our collection of insertional

110

mutants (Pazour et al., 1995, 1998) was screened by real-time PCR with primer pairs specific to

111

NPHP4. In one strain (B1179), no product was amplified. Further analysis by PCR revealed

112

that the mutation, termed nphp4-1, deleted about ~40 kbp around NPHP4 including the entire

113

NPHP4 gene, several predicted genes that have no known association with the flagellum, and a

114

part of DRC3 (Fig. 1A). DRC3 was first identified in our flagellar proteomic study as FAP134

115

(Pazour et al., 2005) and later shown to be a component of the nexin-dynein regulatory complex

116

(Lin et al., 2011). To ensure that the phenotype being analyzed in the studies that follow was not

117

compromised by the absence of DRC3, B1179 was backcrossed twice to a wild-type strain and

118

some of the mutant progeny were then transformed with a DNA fragment containing the DRC3

119

gene (Fig. 1A). One of the resulting transformants, rescued for DRC3, was used in this study as

120

the nphp4 strain.

Journal of Cell Science

121

The nphp4 mutant cells are motile and can phototax (J.A., unpublished observations), but

122

their swimming paths are slightly more erratic than those of wild-type cells (Fig. 1B). To assess

123

this objectively, motility was analyzed using computer-aided sperm analysis (CASA) (Fig. 1C)

124

(Mortimer, 2000). Swimming speeds of wild-type and the nphp4 mutant were similar (VCL,

125

supplementary material Fig. S1B), but linearity was much lower for the mutant (Fig. 1C; see

126

supplementary material Fig. S1A for definition of terms). Transformation of the nphp4 mutant

127

with the wild-type NPHP4 gene to generate strain NPHP4-R almost completely restored linearity

128

(Fig. 1C), confirming that the phenotype is due to the absence of NPHP4. Consistent with the

129

ability to swim nearly normally, steady-state flagella of nphp4 mutant cells are of near normal

130

length and appear to be ultrastructurally normal (Fig. 1D, 2A and J.A., unpublished observations).

131

Following flagellar amputation, the nphp4 mutant regenerated new flagella more slowly than did

132

wild type; normal flagellar regeneration was completely restored in the NPHP4-R strain (Fig.

133

1D). Therefore, NPHP4 is needed to build the flagella with normal kinetics.

134 135

NPHP4 is located in the TZ

136

To determine where NPHP4 is located in C. reinhardtii, we made an antibody to the C-terminus

5

Accepted manuscript Journal of Cell Science

137

of NPHP4. The antibody recognized a band of the expected size in western blots of wild-type

138

but not nphp4 whole cells; the band was restored in the NPHP4-R strain (Fig. 1E). These results

139

confirmed that NPHP4 is missing in the mutant. We also rescued the mutant with constructs

140

expressing a 3xHA tag at either the N- or C-terminus of NPHP4 to generate NPHP4-HAN and

141

NPHP4-HAC cells, respectively; the anti-HA antibody recognized the tagged protein with high

142

specificity in the rescued cells (Fig. 1F). In immunofluorescence microscopy, the anti-NPHP4

143

antibody strongly labeled the bases of the flagella of wild-type cells but not nphp4 cells (Fig. 2A),

144

indicating that this localization is specific to NPHP4. Similarly, the anti-HA antibody localized

145

specifically to the bases of flagella in both NPHP4-HAN and -HAC cells but not in wild-type

146

cells, indicating that the localization is specific to HA-tagged NPHP4 (Fig. 2B).

147

NPHP4 has been localized by immunofluorescence microscopy to primary cilia of

148

MDCK and mouse kidney cells (Mollet et al., 2005; Patil et al., 2012). Therefore, we carefully

149

determined the distribution of NPHP4 in C. reinhardtii flagella vs. cell bodies. When cells are

150

deflagellated, the flagella detach at the junction of the TZ and axoneme proper (Sanders and

151

Salisbury, 1989). NPHP4-HAC cells were deflagellated with dibucaine and the resulting cell

152

bodies and detached flagella were stained with anti-HA; NPHP4-HAC stayed with the cell

153

bodies (Fig. 2C). To confirm this result, whole cells, cell bodies, and isolated flagella were

154

compared by western blotting; all of the NPHP4-HAC remained with the cell body and none

155

could be detected in the isolated flagella (Fig. 2D). The results indicate that C. reinhardtii

156

NPHP4 is located in either the TZ or basal body, but not in the flagella.

157

To further narrow down the location of NPHP4, both nucleo-flagellar apparatuses and

158

whole cells of the NPHP4-HAC strain were labeled with antibodies against the HA-peptide and

159

CEP290, which is located in the TZ (Craige et al., 2010). By both wide-field

160

immunofluorescence microscopy (Fig. 3A, B) and total internal reflection fluorescence/epi-

161

fluorescence structured light microscopy (TESM) (Fig. 3C), the NPHP4-HAC label was distal to

162

the CEP290 label, indicating that NPHP4 is located in the TZ. Analysis of TESM images

163

indicated that the mean peak-to-peak distance between CEP290 and the C-terminus of NPHP4-

164

HAC is about 137 nm (n=130).

165 166

NPHP4 is located at the periphery of the distal TZ

167

To determine the location of NPHP4 within the TZ, we carried out pre-embedding immuno-EM

6

Accepted manuscript Journal of Cell Science

168

of nucleo-flagellar apparatuses from wild-type and NPHP4-HAN and -HAC cells. As expected

169

from the immunofluorescence microscopy results (Fig. 2B), gold particles were associated with

170

the TZ in both NPHP4-HAN and -HAC specimens (Fig. 4 A, B). Most gold particles in

171

longitudinal sections and all particles in cross sections of the TZ were located at the TZ periphery,

172

in close association with remnants of the TZ membrane, which is resistant to treatment with

173

nonionic detergent (Kamiya and Witman, 1984). We observed no gold particles associated with

174

the wild-type TZs, indicating that the immuno-gold labeling was specific to HA-tagged NPHP4.

175

The locations, relative to the proximal end of the TZ, of a total of 50 gold particles from

176

longitudinal sections of both NPHP4-HAN and -HAC specimens were measured and compared

177

to that determined for CEP290 by Craige et al. (2010) (Fig. 4C). Consistent with our

178

immunofluorescence microscopy observations (Fig. 3), the distribution of both NPHP4-HAN

179

and -HAC gold particles had sharp peaks in the distal TZ. The average distance of gold particles

180

from the proximal end of the TZ was 173 ± 37 nm for NPHP4-HAC and 155 ± 31 nm for

181

NPHP4-HAN, compared to 61 ± 38 nm for CEP290; the average distance between NPHP4-HAC

182

and CEP290 gold particles was 111 nm, in good agreement with the separation (~137 nm)

183

measured by TESM.

184 185

NPHP4 and CEP290/NPHP6 localize to the TZ independently of each other

186

To explore if C. reinhardtii NPHP4 can be assembled at the TZ independently from CEP290 and

187

vice versa, we carried out immunofluorescence microscopy of cep290 and nphp4 mutant cells

188

with antibodies against NPHP4 and CEP290, respectively. Localization of NPHP4 appeared to

189

be normal in the cep290 cells, and localization of CEP290 was normal in the nphp4 cells

190

(supplementary material Fig. S2A, B). Therefore, these two proteins are transported to and

191

assemble into the TZ independently of each other.

192 193

NPHP4 does not turnover rapidly

194

We previously reported that CEP290 at the TZ is highly dynamic (Craige et al., 2010). To

195

determine if NPHP4 is similarly dynamic, we mated wild-type gametes to NPHP4-HAN or

196

NPHP4-HAC gametes to form quadriflagellated zygotes in which two TZs contained untagged

197

NPHP4 and two contained HA-tagged NPHP4 in a common cytoplasm. Nucleo-cytoskeletons

198

were prepared at 20, 40, and 60 min after initiation of the mating reaction and labeled with anti-

7

199

HA (Fig. 5). Even after 60 min, only two TZs were labeled by the antibody, indicating that little

200

if any HA-tagged NPHP4 was incorporated into the wild-type TZs over this time period.

201

Moreover, the label was as bright at 60 min as at 20 min, indicating that little if any HA-tagged

202

NPHP4 was replaced with wild-type NPHP4. Therefore, in contrast to CEP290, NPHP4 is static

203

at the TZ.

Journal of Cell Science

Accepted manuscript

204 205

Ultrastructure of the TZ in the nphp4 mutant

206

To determine if loss of NPHP4 caused structural abnormalities in the TZ, we compared the TZs

207

of wild-type and nphp4 mutant cells by EM. In cross sections, the Y-links between the TZ

208

doublet microtubules and membrane, which are disrupted in the cep290 mutant (Craige et al.,

209

2010), were present and appeared normal (Fig. 6A). In longitudinal sections, the wedge-shaped

210

connectors, which extend between the TZ doublets and membrane in wild-type cells, appeared to

211

be missing or collapsed in the nphp4 mutant (Fig. 6B). However, the wedge-shaped structures

212

are similarly altered in TZs of cells lacking IFT protein ift46 (Hou et al., 2007). To clarify if

213

disruption of this structure is correlated with lack of NPHP4, we performed immunofluorescence

214

microscopy of ift46 cells and found that NPHP4 was normally located at the base of their flagella

215

(supplemental material Fig. S2C). Thus, the morphological variation in this structure is not

216

specifically related to loss of NPHP4. Moreover, because the Y-links are present in the nphp4

217

mutant, we can conclude that they and the wedge-shaped structures are not the same.

218

Finally, in ~30% of longitudinal sections of nphp4 flagella, we observed, just distal to the

219

TZ, the apparently ectopic location of dense material resembling the central cylinders of the C.

220

reinhardtii TZ (Fig. 6C).

221 222

Protein composition of the flagellar membrane and matrix is drastically altered by loss of

223

NPHP4

224

To compare the protein compositions of wild-type vs. nphp4 flagella, flagella were isolated and

225

treated with NP-40 to separate proteins into a detergent-soluble membrane-plus-matrix fraction

226

and an axonemal fraction. In SDS-polyacrylamide gels of the axoneme fractions, very few

227

differences were observed (Fig. 7A, B). However, numerous differences were observed in the

228

membrane-plus-matrix fractions. Some proteins were decreased in the nphp4 mutant

229

(arrowheads, Fig. 7A) whereas others were increased (arrows, Fig. 7A). Normal flagellar protein

8

230

composition was restored in NPHP4-R (supplemental material Fig. S3), confirming that the

231

differences were due to loss of NPHP4. Importantly, most differences between the wild-type and

232

nphp4 membrane-plus-matrix fractions were observed in the mass range above 50 kDa; few

233

differences were observed below 50 kDa (Fig. 7B).

Journal of Cell Science

Accepted manuscript

234

To identify the proteins that were decreased or increased in the nphp4 flagella, sets of

235

slices containing the discordant bands were excised from gels of the membrane-plus-matrix

236

fraction of wild-type, nphp4, and NPHP4-R flagella (supplemental material Fig. S3). The slices

237

were then subjected to analysis by mass spectrometry. Multiple proteins were identified in all

238

slices (supplemental material Table S1). To conclude that the amount of a protein was altered in

239

the mutant flagella, we used the following criteria: 1) the number of peptides recovered had to

240

be increased or decreased by more than 10 compared to wild type, or the number of peptides

241

recovered had to be greater than 5 when the protein was not found in the other slice (wild type or

242

nphp4) being compared; and 2) the change was specifically rescued in the NPHP4-R cells.

243

Based on these criteria, the proteins increased or decreased in the nphp4 mutant flagella are listed

244

in Table I and supplemental material Table S1, where blue or red font indicates increased or

245

decreased proteins, respectively. Except for ATPase V1 complex-subunit A, all increased

246

proteins are housekeeping proteins that presumably are abundant in the cell bodies. Valyl- and

247

phenylalanyl-tRNA synthetase, ADP ribosylglycohydrolase, and RNA-binding protein were not

248

found in the C. reinhardtii flagellar proteome (Pazour et al., 2005). All increased proteins are

249

predicted to be larger than 50 kDa (Table I). We conclude that large housekeeping proteins are

250

specifically increased in nphp4 mutant flagella.

251

On the other hand, all decreased proteins other than FAP208 are predicted to be

252

membrane associated: the calcium ATPase is predicted to have a transmembrane domain, and

253

flagellar adenylate kinase, FAP33, FAP295, and FAP12 have N-myristoylation sites.

254

Additionally, when we examined other proteins also predicted to be membrane associated but not

255

meeting the above criteria (these proteins were identified by less than 5 peptides in wild type,

256

were not detected at all in the nphp4 mutant, and were restored in the rescued strain), all were

257

decreased in the mutant flagella. These included the large flagella membrane glycoprotein

258

FMG1-B, a predicted ion channel, another calcium transport ATPase, an ABC transporter, and

259

FAP29, all of which have transmembrane domains; and ankyrin-repeat protein and apyrase,

260

which contain N-myristoylation sites (yellow font in supplemental material Table S1). Nearly all

9

261

decreased proteins were previously identified in the C. reinhardtii flagellar proteome (Pazour et

262

al., 2005). Thus, proteins normally associated with the flagellar membrane are either not

263

transported into or are not retained in the flagella in the absence of NPHP4.

Journal of Cell Science

Accepted manuscript

264

To determine if all flagella-associated membrane proteins are reduced in amount in nphp4

265

flagella, we probed western blots of isolated flagella from wild type and the nphp4 mutant with

266

antibodies against two well-characterized flagellar membrane proteins not identified in our

267

dataset. The flagellar mastigoneme protein (Nakamura et al., 1996) and PKD2 (homologue of

268

human polycystin-2; Huang et al., 2007) were not decreased in the nphp4 flagella (Fig. 7C). We

269

also probed for FMG-1B, which appeared to be decreased in nphp4 flagella in our mass

270

spectrometry analysis; the western blot confirmed that it was reduced in the mutant flagella. We

271

conclude that only a subset of proteins normally associated with the flagellar membrane is

272

decreased in the absence of NPHP4.

273 274

IFT in the nphp4 mutant

275

Several IFT proteins were identified by mass spectrometry and the levels of most of them did not

276

seem to be altered in the mutant flagella (supplemental material Table S1). To confirm this, we

277

examined the amounts of IFT components in isolated flagella of wild type and nphp4 by western

278

blotting. The levels of IFT motors, of IFT proteins of both complex A and B, and of BBS4, a

279

subunit of the BBSome, were unaltered or slightly elevated in the nphp4 mutant flagella

280

(supplemental material Fig. S4A). To more directly assess if IFT is affected by lack of NPHP4,

281

IFT particle movement in steady-state flagella was analyzed (supplemental material Fig. S4B).

282

IFT frequency was not affected whereas IFT velocity was slightly decreased by the absence of

283

NPHP4 (supplemental material Fig. S4C).

284

10

285

DISCUSSION

286

NPHP4 and CEP290 define distinct regions of the TZ

287

In agreement with previous studies in C. elegans and mammals (Winkelbauer et al., 2005;

288

Jauregui et al., 2008; Shiba et al., 2010; Sang et al., 2011; Williams et al., 2011), we found that

289

C. reinhardtii NPHP4 is located in the TZ. However, our studies provide much more precise

290

information on the distribution and localization of NPHP4. Using both immunofluorescence

291

microscopy and immuno-EM, we determined that NPHP4 is located in the distal part of the TZ,

292

distal to CEP290. The NPHP4 and CEP290 peaks observed by TESM were separated by a

Accepted manuscript

293

minimum of ~140 nm, and immuno-EM revealed that there is little overlap between the two

294

proteins. Thus, NPHP4 and CEP290 define distinct structural domains of the TZ (Fig. 7D).

295

Immuno-EM also indicated that the N- and C-termini of NPHP4 are located at the periphery of

296

the TZ, in close association with the TZ membrane. Thus, NPHP4 is well positioned to control

297

entry and/or exit of membrane proteins from the flagellum. This apparent localization of NPHP4

298

to the extreme periphery of the TZ contrasts with that of CEP290, which is distributed

299

throughout the space between the doublet microtubules and the membrane (Craige et al., 2010);

Journal of Cell Science

300

however, we cannot rule out that the middle part of NPHP4 extends into the space between the

301

membrane and microtubules. Finally, our biochemical studies of isolated flagella indicated that

302

there is no detectable NPHP4 in the flagellum proper.

303

Consistent with NPHP4 and CEP290 being located in different regions of the TZ, we

304

found that NPHP4 and CEP290 each assembled normally into the TZ in the absence of the other.

305

Both the localization and assembly results indicate that C. reinhardtii NPHP4 and CEP290 are

306

integrated into different complexes in the TZ. These findings agree with those of Sang et al.

307

(2011), who concluded that mammalian NPHP4 and CEP290 were components of distinct

308

modules based on immunoprecipitations of LAP-tagged proteins.

309

We also observed that NPHP4 at the TZ undergoes little or no turnover during a 1 h

310

period; this is in sharp contrast to CEP290, which is at least 50% replaced in the same time span

311

(Craige et al., 2010). Therefore, NPHP4 is static at the TZ, as might be expected for a structural

312

protein. On the other hand, CEP290 appears to be much more dynamic. Although highly

313

speculative, the results raise the possibility that CEP290 has a role in shuttling IFT or other

314

ciliary proteins into the TZ.

315

11

Accepted manuscript Journal of Cell Science

316

NPHP4 and CEP290 loss have different consequences for TZ structure

317

A highly conserved and defining feature of the TZ is the occurrence of Y-links between the

318

doublet microtubules and membrane as observed in EM cross-sections of various organisms

319

(Gilula and Satir, 1972; Perkins et al., 1986; Ringo, 1967). Craige et al. (2010) observed that the

320

Y-links were missing or disrupted in C. reinhardtii CEP290 mutants. We found that loss of

321

NPHP4, in contrast, did not affect the Y-links. Therefore, NPHP4 is not required for Y-link

322

formation. The same is true in C. elegans, where loss of NPHP4 alone did not cause disruption

323

of the Y-links (Jauregui et al., 2008). However, when mutations in NPHP4 and members of the

324

MKS/B9 module were combined in C. elegans, the Y-links were lost (Williams et al., 2011).

325

Although no specific structural abnormalities were observed in nphp4 TZs, the flagella

326

often contained, distal to the TZ, what appeared to be a partial duplication of the central cylinder

327

that normally occupies the lumen of the TZ of C. reinhardtii. Interestingly, a similar duplication

328

of the central cylinder has been observed in uniflagellar mutants of this organism (Huang et al.,

329

1982; Piasecki et al., 2008). The duplication may occur because the cell detects abnormalities

330

associated with the absence of NPHP4 and responds by moving additional central cylinder

331

precursors into the flagellum.

332 333

A size-dependent barrier to ciliary entry of large soluble housekeeping proteins is breached

334

in nphp4 cells

335

Because of the ease with which C. reinhardtii flagella can be purified, we were able to assess the

336

biochemical consequences of loss of NPHP4 for the flagella. Strikingly, nearly all of the

337

differences observed were in the membrane-plus-matrix fraction, not the axonemal fraction. A

338

large number of proteins were increased in the nphp4 flagella, and, with the possible exception

339

of subunit A of the V1-ATPase, all these proteins were soluble housekeeping proteins. Most of

340

these proteins were not detected at all in the membrane-plus-matrix fraction from wild-type

341

flagella, suggesting that their presence in the nphp4 flagellum resulted from the breakdown of a

342

mechanism that normally keeps them out of the organelle.

343

Importantly, abnormal entry of soluble cytoplasmic proteins into the nphp4 flagellum

344

was observed only for proteins that migrated above 50 kDa in SDS-PAGE, indicating that there

345

is an NPHP4-dependent barrier at the TZ that excludes soluble proteins of ~50 kDa or larger.

346

Consistent with this, Kee et al. (2012) observed that when fluorescently labeled protein A (41

12

Accepted manuscript

347

kDa) and bovine serum albumin (67 kDa) were microinjected into hTERT RPE cells, the former

348

entered the ciliary compartment, whereas the latter did not. Similarly, Breslow et al. (2013),

349

working with permeabilized IMCD3 cells, found that the rate of entry of soluble proteins into

350

the cilium progressively slowed as the proteins increased in size from a Stokes radius of 3.1 nm

351

to 4.3 nm (corresponding to ~30 to ~70 kDa); no entry was observed for larger proteins. Taken

352

together, these results indicate that there is a size-dependent barrier to diffusion of soluble

353

proteins into the cilia of organisms ranging from protists to humans and that this barrier is

354

dependent upon NPHP4 in the TZ.

355

Using a chemically inducible diffusion trap approach, Lin et al. (2013) also obtained

356

evidence for a size-dependent diffusion barrier at the base of primary cilia. However, they

357

observed slow entry of proteins as large as ~600-650 kDa, and suggested that “the steady-state

358

distribution of proteins in cilia is unlikely to be regulated by the diffusion barrier.” Because the

359

differences we observed were in steady-state flagella that were at least 3 hours old, the C.

360

reinhardtii NPHP4-dependent diffusion barrier appears to be more impermeable to the diffusion-

361

based entry of large soluble proteins.

Journal of Cell Science

362 363

NPHP4 is essential to a TZ barrier that retains a subset of membrane proteins in the

364

flagella

365

While levels of large soluble housekeeping proteins increased in nphp4 flagella, levels of many

366

membrane-associated proteins decreased. However, the levels of some other flagellar membrane

367

proteins, such as PKD2 and the mastigoneme protein, were not affected in nphp4 flagella. C.

368

reinhardtii PKD2 is likely to be anchored to the axoneme because most PKD2 peptides were

369

found in axonemal fractions in the flagellar proteome (Pazour et al., 2005) and more than 90% of

370

GFP-tagged PKD2 was stationary in the flagella (Huang et al., 2007). Similarly, 75% of

371

peptides from the mastigoneme protein identified in the flagellar proteome were found in the

372

fraction resistant to detergent, implying that this protein also is anchored to the axoneme (Pazour

373

et al., 2005). These results agree well with reports that PKD2 and an odorant receptor are

374

normally located in C. elegans nphp-4 cilia (Jauregui et al., 2008; Williams et al., 2011). We

375

further observed that the level of FMG-1B was decreased; in the C. reinhardtii flagellar

376

proteome, peptides from FMG-1B were evenly distributed between the membrane and axonemal

377

fractions (Pazour et al., 2005), and at least some of this protein is moved in the plane of the

13

Accepted manuscript Journal of Cell Science

378

flagellar membrane by IFT (Guilford and Bloodgood, 2013). Taken together, these results

379

suggest that a subset of flagellar membrane proteins – specifically those presumed to be mobile

380

in the plane of the membrane – are reduced in nphp4 flagella, whereas those that are anchored to

381

the axoneme exhibit no reduction. Because PKD2 and the mastigoneme protein that bind to the

382

axoneme appear to be delivered normally to the nphp4 flagella, it is possible that most or all

383

flagellar membrane proteins are targeted normally to the flagella, but those that are free to move

384

in the plane of the membrane subsequently diffuse out of the organelle in the absence of an

385

NPHP4-dependent barrier at the TZ.

386

Previous studies have provided evidence for a membrane protein barrier at the base of the

387

cilium that is consistent with our findings. Studies with mammalian cells showed that disruption

388

of members of the MKS/B9 complex, which localizes to the TZ, resulted in a decrease in the

389

amount of several GFP-tagged membrane proteins in primary cilia (Garcia-Gonzalo et al., 2011;

390

Chih et al., 2012). NPHP4 is not part of the MKS/B9 complex, but studies in C. elegans indicate

391

that NPHP-4 and proteins of the MKS/B9 complex genetically interact in functionally redundant

392

pathways important for ciliary development (Williams et al., 2008, 2010, 2011; Huang et al.,

393

2011; Warburton-Pitt et al., 2012). Therefore, NPHP4 and the MKS/B9 proteins likely function

394

together to prevent ciliary membrane proteins from diffusing out of the cilium. NPHP4 is

395

located close to or at the TZ membrane and several of the MKS/B9 proteins are predicted to be

396

transmembrane proteins, so it will be of interest to determine precisely where the MKS/B9

397

proteins are located relative to NPHP4 and CEP290 in the TZ.

398 399

Loss of NPHP4 has mild effects on flagellar motility

400

Interestingly, the only behavioral defect that we observed in the nphp4 mutant was erratic

401

swimming. The lack of a stronger motility defect is consistent with our finding that there is little

402

if any change in the protein composition of the flagellar axoneme, which constitutes the motile

403

machinery. However, changes in intraflagellar Ca2+ are critical for controlling flagellar

404

waveform (Bessen et al., 1980; Kamiya and Witman, 1984), so the erratic swimming that we

405

observed is entirely consistent with the loss of membrane proteins such as the Ca2+-transport

406

ATPases, which are likely involved in maintaining normal intraflagellar Ca2+ levels.

407 408

NPHP4 and CEP290 define different structural and functional domains in the TZ

14

Although nphp4 cells are able to assemble full-length flagella, the kinetics of flagellar

410

regeneration following flagellar amputation is significantly impaired. Presently the reason for

411

this is unclear. IFT frequency is normal and velocity is only slightly reduced in steady-state

412

nphp4 flagella, suggesting that loss of NPHP4 does not directly affect the IFT machinery. These

413

results are generally consistent with those in C. elegans, where mutation of NPHP-4 has modest

414

(Jauregui et al., 2008) or no (Williams et al., 2011) effect on IFT. The slower regeneration

415

kinetics that we observed may reflect reduced IFT cargo loading, impaired flagellar signaling

416

necessary for normal regeneration, or altered intraflagellar ion concentrations (see above

417

section). In any case, the very slight effect of NPHP4 loss on IFT is in contrast to the effect of

418

CEP290 loss, which results in severe IFT abnormalities and greatly impaired flagellar formation

419

(Craige et al., 2010). Therefore, although NPHP4 and CEP290 are both essential to the TZ

420

barrier, CEP290 has an additional function in IFT. We conclude that NPHP4 and CEP290 define

421

different functional as well as structural domains in the TZ.

422

Journal of Cell Science

Accepted manuscript

409

15

423

MATERIALS AND METHODS

Journal of Cell Science

Accepted manuscript

424 425

Strains and growth medium

426

C. reinhardtii wild-type strains 137c (nit1, nit2, mt+) and CC124 (nit1, nit2, mt-) were obtained

427

from the Chlamydomonas Resource Center (University of Minnesota, St. Paul, MN). B1179

428

(nphp4-1, drc3-1, arg7, ARG7, mt+) was generated by insertional mutagenesis of H11 (arg7,

429

mt+) with the pARG7.8 plasmid (Debuchy et al., 1989). The nphp4 strain 79D8 (nphp4-1, drc3,

430

DRC3, mt+), rescued for DRC3, was produced by crossing B1179 sequentially to the wild-type

431

strains CC124 and 137c and then transforming selected progeny with DRC3. NPHP4-R (nphp4-

432

1, NPHP4, drc3, DRC3, mt+) was made by transformation of 79D8 with NPHP4. Two kinds of

433

HA-tagged NPHP4 strains were created: NPHP4-HAN (nphp4-1, NPHP4-HAN, drc3, DRC3,

434

mt+) was generated by transformation of 79D8 with NPHP4-HAN. NPHP4-HAC (nphp4-1,

435

NPHP4-HAC, drc3, DRC3, mt-) was made by transformation of B1179 with NPHP4-HAC

436

followed by two crosses to wild-type strains (CC124 and 137c, sequentially) and transformation

437

of the offspring with DRC3. Cells were grown in M medium I (Sager and Granick, 1953)

438

modified with 2.2 mM KH2PO4 and 1.7 mM K2HPO4 (Witman, 1986) or TAP medium (Gorman

439

and Levine, 1965).

440 441

Real time PCR

442

The deletion site of B1179 was determined by real-time PCR using a thermal cycler (Opticon,

443

MJ Research), the QuantiTect SYBR Green PCR kit (Qiagen), and primers listed in

444

supplemental material Table S2.

445 446

Gene cloning and transformation

447

To clone the full length NPHP4 gene, a 7,159-bp HindIII-SpeI fragment and a 6,357-bp SpeI

448

fragment were cut from BAC clone 18L19 containing NPHP4 and inserted sequentially into

449

pBluescript II SK- (Stratagene). To generate a construct encoding NPHP4-HAN having three

450

consecutive HA peptides (3xHA tag) between the first and second amino acids of NPHP4, a

451

SmaI-NaeI fragment of the p3xHA plasmid (Silflow et al., 2001) was inserted into the AfeI site

452

of NPHP4. To generate a construct encoding NPHP4-HAC having a 3xHA tag between the

453

penultimate and last amino acids, a StuI-SmaI fragment from the p3xHA plasmid was subcloned

16

454

into the PsiI site of NPHP4. To clone DRC3, a 2,225-bp BamHI-SacI fragment and a 5,144-bp

455

SacI fragment were cut from BAC clone 34B19 and inserted into pBluescript II SK-. All BAC

456

clones were purchased from Clemson University Genomics Institute.

Journal of Cell Science

Accepted manuscript

457

All transformations were performed by the glass beads method (Kindle, 1990). For

458

transformation with wild-type and HA-tagged NPHP4, cells were co-transformed with the ble

459

gene as a selectable marker (Stevens et al., 1996). Colonies on TAP medium plates containing

460

zeomycin were randomly selected and screened by immunofluorescence microscopy with anti-

461

NPHP4 or anti-HA antibodies. For transformation with DRC3, the aph7’’ gene was used as the

462

selectable marker (Berthold et al., 2002).

463 464

Analysis of motility

465

Cultures of 137c, 79D8, and NPHP4-R cells at 104-105 cells/ml were diluted 5-10 times with

466

modified M medium for motility analysis. To make a 100-μm deep chamber for observation of

467

the cells, 22 x 22 and 24 x 30 mm No. 1 coverslips were glued to opposite ends of a 25 x 75 mm

468

slide glass so that the two coverslips were separated by about 23 mm. An aliquot (~20 μl) of the

469

cell culture was then placed between the coverslips and covered with a 2X-CEL cover glass

470

(Hamilton Thorne Research). The chamber was inserted into an IVOS sperm analyzer (version

471

12; Hamilton Thorne Research) and cell movement recorded using dark-field optics and a 5x

472

objective.

473 474

Antibodies

475

The last exon of C. reinhardtii NPHP4 was amplified by PCR with primers NPHP4Ex28-N and -

476

C linked to BamHI and SalI sites, respectively. The amplified DNA fragment was cut with those

477

enzymes and put into pMAL-cRI (New England Biolabs) to express the last exon of NPHP4

478

fused to maltose-binding protein (MBP). The fusion proteins were purified with amylose resin

479

and used to produce rabbit polyclonal antibodies (Covance). The NPHP4 antibodies were

480

purified by affinity chromatography with the antigen attached to agarose resin. Bound antibodies

481

were eluted and applied to an agarose column containing MBP to remove anti-MBP antibodies;

482

the flow-through fraction was used in this study. Other antibodies used were anti-βF1-ATPase

483

(Atteia et al., 2006); anti-HA peptide (3F10, Roche); anti-acetylated tubulin (Sigma-Aldrich);

484

anti-α-tubulin (Silflow and Rosenbaum, 1981); anti-CEP290 (Craige et al., 2010); anti-FMG1-B

17

485

(Bloodgood et al., 1986); anti-mastigoneme (Nakamura et al., 1996); anti-PKD2 (Huang et al.,

486

2007); anti-KAP (Mueller et al., 2005); anti-DHC1b (Pazour et al., 1999); anti-IC2 (King et al.,

487

1985); anti-BBS4 (Lechtreck et al., 2009); anti-IFT46 (Hou et al., 2007); and anti-IFT57, anti-

488

IFT81, anti-IFT139, and anti-IFT172 (Cole et al., 1998).

Journal of Cell Science

Accepted manuscript

489 490

Immunofluorescence microscopy

491

Cells were applied for 1-3 min to a coverslip coated with polyethyleneimine, the excess cell

492

suspension removed by blotting, and the coverslip immersed in chilled methanol (-20°C) for 15

493

min. The coverslip was then completely dried before treating with blocking solution (3% fish

494

gelatin, 1% BSA in PBS pH 7.0) for 30 min at RT. Primary antibody diluted in the blocking

495

solution was applied to the sample for overnight at 4°C, followed by 4 washes with PBS.

496

Secondary antibodies diluted in the blocking buffer were then put on the coverslip for 1 h at RT.

497

For double staining, secondary antibodies conjugated to Alexa 488 and 568 or 594 (Life

498

Technologies Corp.) were used. For tri-color labeling, secondary antibodies conjugated to Alexa

499

350, 488, and 594 were used. Finally, the coverslip was rinsed 3 times for 10 min with PBS and

500

mounted on a glass slide with ProLong Gold Antifade Reagent (Life Technologies Corp.).

501

To prepare nucleo-flagellar apparatuses, NPHP4-HAC cells were treated with autolysin,

502

resuspended in ice-cold microtubule-stabilizing (MT) buffer (30 mM Hepes-KOH pH 7.4, 5 mM

503

MgSO4, 15 mM KCl, 2 mM EGTA), and mixed with an equal volume of 1% NP-40

504

(Calbiochem) in MT buffer. The solution was left on ice for 15 min and then mixed with an

505

equal volume of 6% paraformaldehyde in MT buffer. After incubation on ice for 60 min, the

506

nucleo-flagellar apparatuses were collected by centrifugation at 1,100 x g for 10 min. The

507

cytoskeletons were attached to a coverslip coated with poly-L-lysine and dried. The samples

508

were blocked, labeled, and mounted as described above for whole cells.

509

The microscope, camera, and software used to take fluorescence images were described

510

previously (Craige et al., 2010). Gamma adjustment and creation of merged images were done

511

with Photoshop CS2; all images mounted together were processed similarly.

512 513

TESM

514

TESM images were taken using our Biomedical Imaging Group's custom built TESM system

515

(Navaroli et al., 2012). Cells triple labeled with Alexa 488, 594, and 647 and mounted in

18

Accepted manuscript Journal of Cell Science

516

ProLong Gold Antifade Reagent were imaged using three color epi-fluorescence illumination.

517

The pixel size corresponded to 0.083 microns. Z-stacks consisted of 100 images with a z

518

separation of 100 nm. At each z-plane an exposure of 25 ms was taken in each emission

519

bandpass.

520

Tetraspec beads (200 nm) were used to measure chromatic aberration in the microscope

521

system. Point spread functions fitted to the beads in each color revealed a chromatic aberration

522

between the green and the red and far red channels of 1.6 pixels in the x direction. This

523

measured offset was used to correct the experimental data.

524 525

EM

526

Whole cells were pre-fixed in 1% glutaraldehyde in modified M medium for 15 min at RT,

527

collected by centrifugation, resuspended with 1% glutaraldehyde in 100 mM sodium cacodylate-

528

NaOH pH 7.2 (cacodylate buffer), and left for 2 h at RT. Specimens were then processed for

529

ultra-thin sectioning as described previously (Craige et al., 2010).

530

For pre-embedding immunogold EM, nucleo-flagellar apparatuses of 137c, NPHP4-HAN,

531

and NPHP4-HAC cells were prepared as described in “Immunofluorescence microscopy.” After

532

fixation with 3% paraformaldehyde, the samples were washed three times with PBS + 0.02%

533

Tween-20 (PBST) and blocked with 5% skin milk, 4% BSA in PBST for 1 h at RT. The

534

cytoskeletons were then resuspended with the anti-HA antibody diluted 1:400 in PBST and

535

incubated overnight at 4°C. After three washes with PBST, the samples were next incubated

536

with 1% BSA in PBST containing anti-rat IgG conjugated with 10-nm gold particles (Aurion).

537

The cytoskeletons were then washed three times in MT buffer and fixed with 1% OsO4 in MT

538

buffer for 30 min at RT. The pellets were then washed once with MT buffer, rinsed three times

539

with distilled water, and stained with 1% uranyl acetate overnight at 4°C. Dehydration and

540

embedding was performed as for whole cells, except that post-sectioning staining was omitted.

541 542

Generation of gametes and dikaryons

543

To produce gametes, 137c, CC124, NPHP4-HAN, and NPHP4-HAC cells grown to about 105

544

cells/ml in 125 ml of M medium I were collected and resuspended twice with 15 ml of nitrogen-

545

free M medium. The second suspensions were transferred to 250-ml flasks and agitated under

546

constant light overnight. Quadriflagellated zygotes were generated by mixing 137c with

19

547

NPHP4-HAC and CC124 with NPHP4-HAN gametes. At various times after initiation of the

548

mating reaction, aliquots of the mixture were removed and nucleo-flagellar apparatuses prepared

549

as described under “Immunofluorescence microscopy.”

Journal of Cell Science

Accepted manuscript

550 551

Flagella isolation, SDS-PAGE, and western blotting

552

Flagella were isolated by the dibucaine method (Witman, 1986) with the following

553

modifications: DTT was omitted from the HMDS solution. The isolated flagella were

554

resuspended with HMDEK buffer containing 30 mM Hepes pH 7.4, 5 mM MgSO4, 1 mM DTT,

555

1 mM EGTA, and 25 mM KCl. To obtain axonemal and membrane-plus-matrix fractions, NP-40

556

was added to the flagellar suspension to a final concentration of 1% and the suspension left on

557

ice for 10 min. The suspension was then centrifuged (26,890 x g for 10 min at 4°C) and the

558

supernatant collected as the membrane-plus-matrix fraction. The pellet containing axonemes

559

was resuspended in HMDEK buffer. For SDS-PAGE, each fraction was mixed with one-fourth

560

volume of 5xSDS-sample solution (0.3 M Tris-HCl pH 6.8, 5 mM EDTA, 10% SDS, 25%

561

sucrose, and 0.01% bromophenol blue), and DTT added to a final concentration of 50 mM. The

562

mixture was then incubated at 75°C for 20 min. Protein bands were visualized with silver stain

563

(Silver Stain Plus; Bio-Rad). To prepare whole flagella for western blotting, a suspension of

564

isolated flagella was mixed with SDS-sample solution and DTT as described above. For whole

565

cells, cells were collected by centrifugation, lysed with the 5xSDS-sample solution, and

566

incubated at 75°C for 20 min in the presence of 50 mM DTT. Proteins in SDS-polyacrylamide

567

gels were transferred onto Immobilon-P (Millipore). For western blots of whole cells, antibodies

568

were diluted with HIKARI (Nacalai Tescue, Inc.). The chemiluminescence or fluorophore

569

signals from the secondary antibodies were captured with a FluorChem Q equipped with a CCD

570

camera (ProteinSimple).

571 572

Mass spectrometry

573

Protein bands of interest were excised from silver-stained gels and subjected to trypsin digestion

574

as described by Spiess et al. (2011). Mass spectrometry analysis was performed as described by

575

Wagner et al. (2014) with the following modifications: length of LC column was 15 cm and the

576

flow rate of sample injection was 250 nl/min. The resulting spectra were searched against the C.

577

reinhardtii Augustus 5 protein database (Erik Hom, personal communication; based on the Joint

20

578

Genome Institute’s Assembly 4 of the C. reinhardtii genome: http://genome.jgi-

579

psf.org/Chlre4/Chlre4.home.html) using SEQUEST (Bioworks software, v3.3.1; Thermo

580

Electron).

Journal of Cell Science

Accepted manuscript

581 582

Flagella regeneration

583

To determine the kinetics of flagellar regeneration, the pH of mid-log phase cultures of 137c and

584

79D8 cells was lowered to 4.3 with 0.5 M acetic acid. As soon as deflagellation was confirmed

585

by microscopy, the pH was raised to 6.9 with 0.5 M KOH. To fix cells at each time point, a

586

small aliquot of the cell suspension was mixed with an equal volume of M medium containing

587

2% glutaraldehyde. Flagella lengths were measured with the Segmented Line tool in ImageJ

588

(National Institutes of Health) and mean values at each point were obtained from 50 flagella on

589

25 cells.

590 591

Observation of IFT

592

137c and nphp4 cells at about 104 cells/ml were used to observe IFT as previously described

593

(Craige et al., 2010). Kymographs were made by ImageJ with the MultipleKymograph plugin.

594

21

595

ACKNOWLEDGEMENTS

596

We are grateful to all those who generously provided antibodies, to the University of

597

Massachusetts Medical School (UMMS) Core EM Facility for expert help with EM, and to the

598

Vermont Genetics Network (VGN) Proteomics Facility for mass spectrometry. We thank Eric

599

Coleman Johnson for identifying the nphp4 mutant. We also thank Erik Hom (Harvard

600

University) for providing C. reinhardtii Augustus 10 gene models, and Branch Craige (UMMS)

601

for critical discussion.

Accepted manuscript

603

609

Funding

Journal of Cell Science

602

610

This research was supported by NIH grants GM060992 (to GJ) and GM030626 (to GW), by the

611

Robert W. Booth Endowment at UMMS (to GW), and by NIGMS Institutional Development

612

Award (IDeA) P20 GM103449 (to the VGN).

Author contributions

604

J.A. designed and performed most experiments. S.T. did the initial characterization of the nphp4

605

mutant. C.S., K.D.B., and K.E.F. acquired data with TESM. K.F.L. assisted with some

606

experiments. J.A. and G.B.W. prepared the manuscript. G.J.P and G.B.W. supervised and

607

funded this study and helped interpret the data.

608

613 614

22

615

REFERENCES

616

Atteia, A., van Lis, R., Gelius-Dietrich, G., Adrait, A., Garin, J., Joyard, J., Rolland, N. and

617

Martin, W. (2006). Pyruvate formate-lyase and a novel route of eukaryotic ATP

618

synthesis in Chlamydomonas mitochondria. J Biol Chem. 281, 9909-9918.

619

Berthold, P., Schmitt, R. and Mages, W. (2002). An engineered Streptomyces hygroscopicus

621

reinhardtii. Protist. 153, 401-412.

Accepted manuscript

aph 7" gene mediates dominant resistance against hygromycin B in Chlamydomonas

622

Journal of Cell Science

620

628

Antignac, C., Saunier, S. and Schneider-Maunoury, S. (2011). Control of the Wnt

629

pathways by nephrocystin-4 is required for morphogenesis of the zebrafish pronephros.

630

Hum Mol Genet. 20, 2611-2627.

623 624

Bessen, M., Fay, R. B. and Witman, G. B. (1980). Calcium control of waveform in isolated flagellar axonemes of Chlamydomonas. J Cell Biol. 86, 446-455. Bloodgood, R. A., Woodward, M. P. and Salomonsky, N. L. (1986). Redistribution and

625

shedding of flagellar membrane glycoproteins visualized using an anti-carbohydrate

626

monoclonal antibody and concanavalin A. J Cell Biol. 102, 1797-1812.

627

631

Burcklé, C., Gaude, H. M., Vesque, C., Silbermann, F., Salomon, R., Jeanpierre, C.,

Breslow, D. K., Koslover, E. F., Seydel, F., Spakowitz, A. J. and Nachury, M. V. (2013). An

632

in vitro assay for entry into cilia reveals unique properties of the soluble diffusion barrier.

633

J Cell Biol. 203, 129-147.

634

Chih, B., Liu, P., Chinn, Y., Chalouni, C., Komuves, L. G., Hass, P. E., Sandoval, W. and

635

Peterson A. S. (2012). A ciliopathy complex at the TZ protects the cilia as a privileged

636

membrane domain. Nat Cell Biol. 14, 61-72.

637

Cole, D. G., Diener, D. R., Himelblau, A. L., Beech, P. L., Fuster, J. C. and Rosenbaum, J.

638

L. (1998). Chlamydomonas kinesin-II-dependent intraflagellar transport (IFT): IFT

639

particles contain proteins required for ciliary assembly in Caenorhabditis elegans sensory

640

neurons. J Cell Biol. 141, 993-1008.

23

641 642

Witman, G. B. (2010). CEP290 tethers flagellar TZ microtubules to the membrane and

643

regulates flagellar protein content. J Cell Biol. 190, 927-940.

644 645

Accepted manuscript

646

Journal of Cell Science

Craige, B., Tsao, C. C., Diener, D. R., Hou, Y., Lechtreck, K. F., Rosenbaum, J. L. and

Czarnecki, P. G. and Shah, J. V. (2012). The ciliary transition zone: from morphology and molecules to medicine. Trends Cell Biol. 22, 201-210. Debuchy, R., Purton, S. and Rochaix, J. D. (1989). The argininosuccinate lyase gene of

647

Chlamydomonas reinhardtii: an important tool for nuclear transformation and for

648

correlating the genetic and molecular maps of the ARG7 locus. EMBO J. 8, 2803-2809.

649

Failler, M., Gee, H. Y., Krug, P., Joo, K., Halbritter, J., Belkacem, L., Filhol, E., Porath, J.

650

D., Braun, D. A., Schueler, M. et al. (2014). Mutations of CEP83 cause infantile

651

nephronophthisis and intellectual disability. Am J Hum Genet. 94, 905-914.

652 653 654

Fisch, C. and Dupuis-Williams, P. (2011). Ultrastructure of cilia and flagella – back to the future! Biol. Cell 103, 249-270. Garcia-Gonzalo, F. R., Corbit, K. C., Sirerol-Piquer, M. S., Ramaswami, G., Otto, E. A.,

655

Noriega, T. R., Seol, A. D., Robinson, J. F., Bennett, C. L., Josifova, D. J. et al.

656

(2011). A TZ complex regulates mammalian ciliogenesis and ciliary membrane

657

composition. Nat Genet. 43, 776-784.

658 659 660

Gilula, N. B. and Satir, P. (1972). The ciliary necklace. A ciliary membrane specialization. J Cell Biol. 53, 494-509. Gorman, D. S. and Levine, R. P. (1965). Cytochrome f and plastocyanin: their sequence in the

661

photosynthetic electron transport chain of Chlamydomonas reinhardi. Proc Natl Acad Sci

662

U S A. 54, 1665-1669.

663 664

Guilford, W. H. and Bloodgood, R. A. (2013). Laser trap measurements of flagellar membrane motility. Methods Enzymol. 525, 85-107.

665

24

666 667

B., Attanasio, M., O'Toole, J. F. et al. (2005). Mutational analysis of the NPHP4 gene

668

in 250 patients with nephronophthisis. Hum Mutat. 25, 411.

Accepted manuscript

669

Journal of Cell Science

Hoefele, J., Sudbrak, R., Reinhardt, R., Lehrack, S., Hennig, S., Imm, A., Muerb, U., Utsch,

Hou, Y., Qin, H., Follit, J. A., Pazour, G. J., Rosenbaum, J. L. and Witman, G. B. (2007).

670

Functional analysis of an individual IFT protein: IFT46 is required for transport of outer

671

dynein arms into flagella. J Cell Biol. 176, 653-665.

672

Huang, B., Ramanis, Z., Dutcher, S. K. and Luck, D. J. (1982). Uniflagellar mutants of

673

Chlamydomonas: evidence for the role of basal bodies in transmission of positional

674

information. Cell. 29, 745-753.

675

Huang, K., Diener, D. R., Mitchell, A., Pazour, G. J., Witman, G. B. and Rosenbaum, J. L.

676

(2007). Function and dynamics of PKD2 in Chlamydomonas reinhardtii flagella. J Cell

677

Biol. 179, 501-514.

678

Huang, L., Szymanska, K., Jensen, V. L., Janecke, A. R., Innes, A. M., Davis, E. E., Frosk,

679

P., Li, C., Willer, J. R., Chodirker, B. N. et al. (2011). TMEM237 is mutated in

680

individuals with a Joubert syndrome related disorder and expands the role of the TMEM

681

family at the ciliary transition zone. Am J Hum Genet. 89, 713-730.Ishikawa, H. and

682

Marshall, M. F. (2011). Ciliogenesis: building the cell’s antenna. Nat Rev Mol Cell Biol.

683

12, 222-234.

684

Jauregui, A. R., Nguyen, K. C., Hall, D. H. and Barr, M. M. (2008). The Caenorhabditis

685

elegans nephrocystins act as global modifiers of cilium structure. J Cell Biol. 180, 973-

686

988.

687

Kamiya, R. and Witman, G. B. (1984). Submicromolar levels of calcium control the balance of

688

beating between the two flagella in demembranated models of Chlamydomonas. J Cell

689

Biol. 98, 97-107.

690

Kee, H. L., Dishinger, J. F., Blasius, T. L., Liu, C. J., Margolis, B. and Verhey, K. J. (2012).

691

A size-exclusion permeability barrier and nucleoporins characterize a ciliary pore

692

complex that regulates transport into cilia. Nat Cell Biol. 14, 431-437.

25

693 694

Proteomic analysis of isolated Chlamydomonas centrioles reveals orthologs of ciliary-

695

disease genes. Curr Biol. 15, 1090-1098.

696 697

Accepted manuscript

698

Journal of Cell Science

Keller, L. C., Romijn, E. P., Zamora, I., Yates, J. R., 3rd and Marshall, W. F. (2005).

Kindle, K. L. (1990). High-frequency nuclear transformation of Chlamydomonas reinhardtii. Proc Natl Acad Sci U S A. 87, 1228-1232. King, S. M., Otter, T. and Witman, G. B. (1985). Characterization of monoclonal antibodies

699

against Chlamydomonas flagellar dyneins by high-resolution protein blotting. Proc Natl

700

Acad Sci U S A. 82, 4717-4721.

701

Lechtreck, K. F., Johnson, E. C., Sakai, T., Cochran, D., Ballif, B. A., Rush, J., Pazour, G.

702

J., Ikebe, M. and Witman, G. B. (2009). The Chlamydomonas reinhardtii BBSome is

703

an IFT cargo required for export of specific signaling proteins from flagella. J Cell Biol.

704

187, 1117-1132.

705

Lin, J., Tritschler, D., Song, K., Barber, C. F., Cobb, J. S., Porter, M. E. and Nicastro, D.

706

(2011). Building blocks of the nexin-dynein regulatory complex in Chlamydomonas

707

flagella. J Biol Chem. 286, 29175-29191.

708

Lin, Y. C., Niewiadomski, P., Lin, B., Nakamura, H., Phua, S. C., Jiao, J., Levchenko, A.,

709

Inoue, T., Rohatgi, R. and Inoue, T. (2013). Chemically inducible diffusion trap at cilia

710

reveals molecular sieve–like barrier. Nat Chem Biol. 9, 437-443.

711

Mollet, G., Salomon, R., Gribouval, O., Silbermann, F., Bacq, D., Landthaler, G., Milford,

712

D., Nayir, A., Rizzoni, G., Antignac, C. et al. (2002). The gene mutated in juvenile

713

nephronophthisis type 4 encodes a novel protein that interacts with nephrocystin. Nat

714

Genet. 32, 300-305.

715

Mollet, G., Silbermann, F., Delous, M., Salomon, R., Antignac, C. and Saunier, S. (2005).

716

Characterization of the nephrocystin/nephrocystin-4 complex and subcellular localization

717

of nephrocystin-4 to primary cilia and centrosomes. Hum Mol Genet. 14, 645-656.

718

Mortimer, S. T. (2000). CASA--practical aspects. J Androl. 21, 515-524.

26

719

Mueller, J., Perrone, C. A., Bower, R., Cole, D. G. and Porter, M. E. (2005). The FLA3 KAP

720

subunit is required for localization of kinesin-2 to the site of flagellar assembly and

721

processive anterograde intraflagellar transport. Mol Biol Cell. 16, 1341-1354.

722 723

Assembly and function of Chlamydomonas flagellar mastigonemes as probed with a

724

monoclonal antibody. J Cell Sci. 109 (Pt 1), 57-62.

Accepted manuscript

725

Journal of Cell Science

Nakamura, S., Tanaka, G., Maeda, T., Kamiya, R., Matsunaga, T. and Nikaido, O. (1996).

Navaroli, D. M., Bellvé, K. D., Standley, C., Lifshitz, L. M., Cardia, J., Lambright, D.,

726

Leonard, D., Fogarty, K. E. and Corvera, S. (2012). Rabenosyn-5 defines the fate of

727

the transferrin receptor following clathrin-mediated endocytosis. Proc Natl Acad Sci U S

728

A. 109, E471-E480.

729

Otto, E., Hoefele, J., Ruf, R., Mueller, A. M., Hiller, K. S., Wolf, M. T., Schuermann, M. J.,

730

Becker, A., Birkenhager, R., Sudbrak, R. et al. (2002). A gene mutated in

731

nephronophthisis and retinitis pigmentosa encodes a novel protein, nephroretinin,

732

conserved in evolution. Am J Hum Genet. 71, 1161-1167.

733

Patil, H., Tserentsoodol, N., Saha, A., Hao, Y., Webb, M. and Ferreira, P. A. (2012).

734

Selective loss of RPGRIP1-dependent ciliary targeting of NPHP4, RPGR and SDCCAG8

735

underlies the degeneration of phtoreceptor neurons. Cell Death Dis. 3, e355.

736

Pazour, G. J., Sineshchekov, O. A. and Witman, G. B. (1995). Mutational analysis of the

737 738

phototransduction pathway of Chlamydomonas reinhardtii. J Cell Biol. 131, 427-440. Pazour, G. J., Wilkerson, C. G. and Witman, G. B. (1998). A dynein light chain is essential

739

for the retrograde particle movement of intraflagellar transport (IFT). J Cell Biol. 141,

740

979-992.

741 742 743 744

Pazour, G. J., Dickert, B. L. and Witman, G. B. (1999). The DHC1b (DHC2) isoform of cytoplasmic dynein is required for flagellar assembly. J Cell Biol. 144, 473-481. Pazour, G. J., Agrin, N., Leszyk, J. and Witman, G. B. (2005). Proteomic analysis of a eukaryotic cilium. J Cell Biol. 170, 103-113.

27

745 746 747

Journal of Cell Science

Accepted manuscript

748

Pedersen, L. B. and Rosenbaum, J. L. (2008). Intraflagellar transport (IFT) role in ciliary assembly, resorption and signalling. Curr Top Dev Biol. 85, 23-61. Perkins, L. A., Hedgecock, E. M., Thomson, J. N. and Culotti, J. G. (1986). Mutant sensory cilia in the nematode Caenorhabditis elegans. Dev Biol. 117, 456-487.

749

Piasecki, B. P., LaVoie, M., Tam, L. W., Lefebvre, P. A. and Silflow, C. D. (2008). The Uni2

750

phosphoprotein is a cell cycle regulated component of the basal body maturation pathway

751

in Chlamydomonas reinhardtii. Mol Biol Cell. 19, 262-273.

752

Reiter, J. F., Blacque, O. E. and Leroux, M. R. (2012). The base of the cilium: roles for

753

transition fibres and the transition zone in ciliary formation, maintenance and

754

compartmentalization. EMBO Rep.13, 608-618.

755

Renkema, K. Y., Stokman, M. F., Giles, R. H. and Knoers, N. V. (2014). Next-generation

756

sequencing for research and diagnostics in kidney disease. Nat Rev Nephrol. June 10, in

757

press.

758 759 760

Ringo, D. L. (1967). Flagellar motion and fine structure of the flagellar apparatus in Chlamydomonas. J Cell Biol. 33, 543-571. Roepman, R., Letteboer, S. J., Arts, H. H., van Beersum, S. E., Lu, X., Krieger, E., Ferreira,

761

P. A. and Cremers, F. P. (2005). Interaction of nephrocystin-4 and RPGRIP1 is

762

disrupted by nephronophthisis or Leber congenital amaurosis-associated mutations. Proc

763

Natl Acad Sci U S A. 102, 18520-18525.

764 765 766 767

Rosenbaum, J. L. and Witman, G. B. (2002). Intraflagellar transport. Nat Rev Mol Cell Biol. 3, 813-825. Sager, R. and Granick, S. (1953). Nutritional studies with Chlamydomonas reinhardi. Ann N Y Acad Sci. 56, 831-838.

768

Sanders, M. A. and Salisbury, J. L. (1989). Centrin-mediated microtubule severing during

769

flagellar excision in Chlamydomonas reinhardtii. J Cell Biol. 108, 1751-1760.

28

Journal of Cell Science

Accepted manuscript

770

Sang, L., Miller, J. J., Corbit, K. C., Giles, R. H., Brauer, M. J., Otto, E. A., Baye, L. M.,

771

Wen, X., Scales, S. J., Kwong, M. et al. (2011). Mapping the NPHP-JBTS-MKS

772

protein network reveals ciliopathy disease genes and pathways. Cell. 145, 513-528.

773

Schuermann, M. J., Otto, E., Becker, A., Saar, K., Ruschendorf, F., Polak, B. C., Ala-Mello,

774

S., Hoefele, J., Wiedensohler, A., Haller, M. et al. (2002). Mapping of gene loci for

775

nephronophthisis type 4 and Senior-Loken syndrome, to chromosome 1p36. Am J Hum

776

Genet. 70, 1240-1246.

777

Shiba, D., Manning, D. K., Koga, H., Beier, D. R. and Yokoyama, T. (2010). Inv acts as a

778

molecular anchor for Nphp3 and Nek8 in the proximal segment of primary cilia.

779

Cytoskeleton. 67, 112-119.

780 781 782

Shiba, D. and Yokoyama, T. (2012). The ciliary transitional zone and nephrocystins. Differentiation. 83, S91-S96. Silflow, C. D. and Rosenbaum, J. L. (1981). Multiple alpha- and beta-tubulin genes in

783

Chlamydomonas and regulation of tubulin mRNA levels after deflagellation. Cell. 24, 81-

784

88.

785

Silflow, C. D., LaVoie, M., Tam, L. W., Tousey, S., Sanders, M., Wu, W., Borodovsky, M.

786

and Lefebvre, P. A. (2001). The Vfl1 protein in Chlamydomonas localizes in a

787

rotationally asymmetric pattern at the distal ends of the basal bodies. J Cell Biol. 153, 63-

788

74.

789

Slanchev, K., Putz, M., Schmitt, A., Kramer-Zucker, A. and Walz, G. (2011). Nephrocystin-

790

4 is required for pronephric duct-dependent cloaca formation in zebrafish. Hum Mol

791

Genet. 20, 3119-3128.

792

Spiess, P. C., Deng, B., Hondal, R. J., Matthews, D. E. and van der Vliet, A. (2011).

793

Proteomic profiling of acrolein adducts in human lung epithelial cells. Journal of

794

proteomics. 74, 2380-2394.

795 796

Stevens, D. R., Rochaix, J. D. and Purton, S. (1996). The bacterial phleomycin resistance gene ble as a dominant selectable marker in Chlamydomonas. Mol Gen Genet. 251, 23-30.

29

797 798

Brooks, E. M., Platz, J. J., Khalpey, Z. I., Hoganson, D. M. et al. (2014). Three-

799

dimensional scaffolds of acellular human and porcine lungs for high throughput studies

800

of lung disease and regeneration. Biomaterials. 35, 2664-2679

Accepted manuscript

801

Journal of Cell Science

Wagner, D. E., Bonenfant, N. R., Sokocevic, D., DeSarno, M. J., Borg, Z. D., Parsons, C. S.,

Warburton-Pitt, S. R., Jauregui, A. R., Li, C., Wang, J., Leroux, M. R. and Barr, M. M.

802

(2012). Ciliogenesis in Caenorhabditis elegans requires genetic interactions between

803

ciliary middle segment localized NPHP-2 (inversin) and transition zoneassociated

804

proteins. J Cell Sci. 125, 2592-2603.

805

Williams, C. L., Winkelbauer, M. E., Schafer, J. C., Michaud, E. J. and Yoder, B. K. (2008).

806

Functional redundancy of the B9 proteins and nephrocystins in Caenorhabditis elegans

807

ciliogenesis. Mol Cell Biol. 19, 2154-2168.

808

Williams, C. L., Masyukova, S. V. and Yoder, B. K. (2010). Normal ciliogenesis requires

809

synergy between the cystic kidney disease genes MKS-3 and NPHP-4. J Am Soc Nephrol.

810

21, 782-793.

811

Williams, C. L., Li, C., Kida, K., Inglis, P. N., Mohan, S., Semenec, L., Bialas, N. J., Stupay,

812

R. M., Chen, N., Blacque, O. E. et al. (2011). MKS and NPHP modules cooperate to

813

establish basal body/TZ membrane associations and ciliary gate function during

814

ciliogenesis. J Cell Biol. 192, 1023-1041.

815

Winkelbauer, M. E., Schafer, J. C., Haycraft, C. J., Swoboda, P. and Yoder, B. K. (2005).

816

The C. elegans homologs of nephrocystin-1 and nephrocystin-4 are cilia TZ proteins

817

involved in chemosensory perception. J Cell Sci. 118, 5575-5587.

818 819

Witman, G. B. (1986). Isolation of Chlamydomonas flagella and flagellar axonemes. Methods Enzymol. 134, 280-290.

820

Wolf, M. T. and Hildebrandt, F. (2011). Nephronophthisis. Pediatr Nephrol. 26, 181-194.

821

Wolf, M. T., Lee, J., Panther, F., Otto, E. A., Guan, K. L. and Hildebrandt, F. (2005).

822

Expression and phenotype analysis of the nephrocystin-1 and nephrocystin-4 homologs in

823

Caenorhabditis elegans. J Am Soc Nephrol. 16, 676-687.

30

824

Won, J., Marin de Evsikova, C., Smith, R. S., Hicks, W. L., Edwards, M. M., Longo-Guess,

825

C., Li, T., Naggert, J. K. and Nishina, P. M. (2011). NPHP4 is necessary for normal

826

photoreceptor ribbon synapse maintenance and outer segment formation, and for sperm

827

development. Hum Mol Genet. 20, 482-496.

Journal of Cell Science

Accepted manuscript

828

31

829

FIGURE LEGENDS

830

Figure 1. Characterization of C. reinhardtii nphp4 mutant. (A) Map of C. reinhardtii

831

genome near the NPHP4 locus. Numbers above each locus correspond to gene IDs in

832

Phytozome version 9.1 (http://www.phytozome.net). Arrows indicate the positions of PCR

833

products used to delimit the deleted region; plus and minus marks indicate whether the PCR

834

products were amplified. Genome fragments from bacterial artificial chromosome (BAC) clones

835

used for knock-in of full-length NPHP4 and DRC3 are indicated by white rectangles. (B)

836

Representative images showing swimming paths of wild-type and nphp4 cells. White dots show

Accepted manuscript

837

cells where CASA began to monitor their tracks; green lines are the swimming paths analyzed.

838

Blue tracks were not analyzed because the cells swam outside the microscope field before

839

recording was completed. Cells marked with red dots were immobilized by attachment to the

840

coverslip. The swimming paths of nphp4 cells are more erratic than those of wild-type cells. (C)

841

Means ± s.d. of linearity and CASA histogram summarizing distribution of linearity in

842

swimming paths of populations of wild-type, nphp4, and NPHP4-R cells (see supplementary

843

material Fig. S1). For each strain, the population was calculated from the sum of values from a

Journal of Cell Science

844

total of 10 fields in each of 5 independent experiments. Statistical significance was determined

845

by the Tukey-Kramer method: **,

NPHP4 controls ciliary trafficking of membrane proteins and large soluble proteins at the transition zone.

The protein nephrocystin-4 (NPHP4) is widespread in ciliated organisms, and defects in NPHP4 cause nephronophthisis and blindness in humans. To learn ...
2MB Sizes 0 Downloads 6 Views