当前位置: > 正文
编号:11374374
Polyglycylation of Tubulin Is Essential and Affects Cell Motility and
http://www.100md.com 《细胞学杂志》
     a Department of Cellular Biology, University of Georgia, Athens, Georgia 30602-2607

    b Department of Biology, University of Rochester, Rochester, New York 14627

    c Laboratoire de Biologie Cellulaire 4, CNRS UPRES-A 8080, Université Paris XI, 91405, Orsay Cedex, France

    Correspondence to: Jacek Gaertig, Department of Cellular Biology, 724 Biological Sciences Building, University of Georgia, Athens, GA 30602-2607. Tel:(706) 542-3409 Fax:(706) 542-4271 E-mail:jgaertig@cb.uga.edu.

    Abstract

    We analyzed the role of tubulin polyglycylation in Tetrahymena thermophila using in vivo mutagenesis and immunochemical analysis with modification-specific antibodies. Three and five polyglycylation sites were identified at glutamic acids near the COOH termini of - and ?-tubulin, respectively. Mutants lacking all polyglycylation sites on -tubulin have normal phenotype, whereas similar sites on ?-tubulin are essential. A viable mutant with three mutated sites in ?-tubulin showed reduced tubulin glycylation, slow growth and motility, and defects in cytokinesis. Cells in which all five polyglycylation sites on ?-tubulin were mutated were viable if they were cotransformed with an -tubulin gene whose COOH terminus was replaced by the wild-type COOH terminus of ?-tubulin. In this double mutant, ?-tubulin lacked detectable polyglycylation, while the -? tubulin chimera was hyperglycylated compared with -tubulin in wild-type cells. Thus, the essential function of polyglycylation of the COOH terminus of ?-tubulin can be transferred to -tubulin, indicating it is the total amount of polyglycylation on both - and ?-tubulin that is essential for survival.

    Key Words: motor proteins, microtubules, cilia, cytoskeleton, motility

    Introduction

    Microtubules (MTs)1 are ubiquitous eukaryotic cytoskeletal filaments consisting of linear polymers of noncovalently linked heterodimers of - and ?-tubulin. They form the structural framework of organelles that are essential for maintaining shape, and play major roles in cell movement, intracellular transport, and cell division. Different MT systems (e.g., cytoplasmic network, axonemes, centrioles, and mitotic spindle) contain MTs that have distinct lengths, spatial arrangements, and dynamics. Little is known about the mechanisms that result in the correct location, assembly, and function of distinct microtubules in a cell.

    It seems reasonable to expect that the functional diversity of microtubules reflects underlying compositional differences among them. However, most cell types do not express enough tubulin isotypes to account for all microtubule diversity (Silflow 1991 ). Ciliates, in particular, have a small number of tubulin isotypes while exhibiting diversity of microtubules comparable to that found in an entire multicellular organism (Conzelmann and Helftenbein 1987 ; Gaertig et al. 1993 ).

    In most eukaryotes, posttranslational tubulin modifications (PTMs; for reviews see MacRae 1997 ; Luduena 1998 ) also produce heterogeneity among - and ?-tubulins. These modifications include the following: -tubulin acetylation (L'Hernault and Rosenbaum 1985 ), removal of the COOH-terminal tyrosine of -tubulin (Argarana et al. 1978 ), phosphorylation of ?-tubulin (Eipper 1972 ), palmitoylation of -tubulin (Caron 1997 ), polyglutamylation (Edde et al. 1990 ; Alexander et al. 1991 ; Redeker et al. 1992 ; Rudiger et al. 1992 ; Mary et al. 1994 ), and polyglycylation (Redeker et al. 1994 ) of - and ?-tubulin.

    Polyglycylation is a recently identified PTM, so far unique to tubulin, that adds up to 34 glycine residues onto the -COOH groups of specific glutamate residues (Es) of - and ?-tubulin in Paramecium (Redeker et al. 1994 ). Polyglycylation occurs on several clustered Es of ?-tubulin of Paramecium including E437, E438, E439, and E441 (Vinh et al. 1999 ). This PTM has been found in diverse species (Rudiger et al. 1995 ; Bre et al. 1996 ; Mary et al. 1996 ; Multigner et al. 1996 ; Weber et al. 1996 ), but only in cell types that have either cilia or flagella (discussed in Levilliers et al. 1995 ; Bre et al. 1998 ). mAbs specific for polyglycylated tubulins inhibited the motility of reactivated sea urchin spermatozoa (Bre et al. 1996 ), suggesting that this PTM plays a role in regulation of ciliary dynein.

    We have been developing the ciliated protozoan, Tetrahymena thermophila, as a model for studying the in vivo function of PTMs. Tetrahymena and Paramecium belong to the same class of ciliates (Baroin-Tourancheau et al. 1992 ) and have similar cytoskeletal organizations (Fleury et al. 1992 ). T. thermophila maintains at least 17 distinct microtubule structures, but expresses only one type of -tubulin, one major and one highly divergent minor ?-tubulin proteins encoded by a single - and three ?-tubulin genes (Gaertig et al. 1993 ; McGrath et al. 1994 ; Li, B., and M.A. Gorovsky, unpublished results). However, Tetrahymena tubulins occur in multiple isoforms generated by various PTMs (Suprenant et al. 1985 ; Gaertig et al. 1995 ). The COOH termini of Tetrahymena - and ?-tubulin are similar to the conserved COOH termini of other axonemal tubulins, and contain several possible sites of polyglycylation. Here, we describe genetic analyses of the polyglycylatable sites of Tetrahymena - and ?-tubulin. We show that cells need polyglycylation sites on ?-tubulin for survival, whereas similar sites on -tubulin are dispensable. However, a lethal polyglycylation site mutation on ?-tubulin could be rescued if the COOH-terminal domain of -tubulin was replaced with the wild-type COOH terminus of ?-tubulin. Thus, polyglycylation of - and ?-tubulin appears to have redundant functions and the total amount of polyglycylation on both subunits appears to be critical for cell survival.

    Materials and Methods

    Cell Culture

    T. thermophila cells were grown in SPPA (1% proteose peptone, 0.2% glucose, 0.1% yeast extract, 0.003% EDTA·ferric sodium salt, 100 U/ml penicillin, 100 μg/ml streptomycin, 0.25 μg/ml fungizone) at 30°C with shaking.

    Germline Transformation and Construction of BTU Double Knockout Heterokaryons

    To disrupt the T. thermophila BTU2 gene, we constructed the pTBS plasmid with the coding sequence of the 3-kb HindIII fragment of BTU2 (Gaertig et al. 1993 ) replaced by the blasticidin S (bs) resistance gene, bsr, (Sutoh 1993 ) which is controlled by the HHF1 gene promoter (Gaertig et al. 1994a ). For disruption of BTU1, we used the pHAB1 plasmid containing a BTU1 fragment whose coding sequence was replaced by the neo gene under control of the HHF1 promoter. To disrupt genes in the germline micronucleus, pTBS or pHAB1 DNA was purified using the Plasmid Maxi kit (Qiagen). The pTBS plasmid was linearized with EcoRI and SalI to release the btu2::bsr1 insert, whereas the pHAB1 plasmid was linearized with SacI and SalI to release the btu1::neo1 insert. Biolistic germline transformation was done as previously described (Cassidy-Hanley et al. 1997 ), except that we used gold particles (0.6 μm; Bio-Rad Laboratories) instead of tungsten. The btu2::bsr1 transformants were selected at 60 μg/ml blasticidin S (ICN), whereas the btu1::neo1 transformants were selected with 120 μg/ml paromomycin (pm; Sigma Chemical Co.) in SPPA. Transformants were confirmed to be heterozygous germline knockouts as described previously (Cassidy-Hanley et al. 1997 ). A heterozygous clone for the btu2::bsr1/BTU2 was crossed to a strain heterozygous for the btu1::neo1/BTU1 gene in the micronucleus. Double heterozygotes from this cross were mated to a B*VII strain (Orias and Bruns 1976 ) to obtain cells with micronuclei homozygous for both disrupted alleles. Two exconjugant clones of different mating types (DB2A and DB6B) were identified as homozygotes for both disrupted BTU genes based on the appearance of bs-r and pm-r progeny in an outcross.

    Introduction of Mutated Tubulin Genes by Rescue of Mating Knockout Heterokaryons

    Plasmid pTUB100E3-PvuII contains the 3.2-kb HindIII genomic fragment of the -tubulin gene of T. thermophila, ATU1 (Hai and Gorovsky 1997 ). Derivatives of pTUB100E3-PvuII containing desired mutations were made by site-directed mutagenesis (Kunkel 1985 ). The same approach was used with plasmid pBTU1 (Gaertig et al. 1993 ) to construct mutations of the T. thermophila BTU1 gene encoding ?-tubulin. The ATU1 genes (linearized with HindIII) were used to rescue mating heterokaryon strains lacking functional -tubulin genes in the micronucleus (AAKO2 and AAKO5), and transformed cells were selected with 120 μg/ml pm as previously described (Hai and Gorovsky 1997 ). Introduction of mutated BTU1 genes (linearized with XbaI and HindIII) was performed by rescue transformation of mating heterokaryon strains DB2A and DB6B, and transformed progeny were selected with 60 μg/ml of bs in SPPA.

    Isolation of Tetrahymena Genomic DNA and Southern Blotting

    Late log phase Tetrahymena cells (6–8 x 105 cells/ml) were starved overnight in 10 mM Tris, pH 7.5, at 30°C. Total genomic DNA was isolated from 50 ml of starved cells and Southern blots were prepared as described (Gaertig et al., 1994). Radiolabeled fragments of the 3' flanking sequence of either BTU1 or BTU2 gene were used as probes.

    Isolation of Cilia

    A 1-liter culture of Tetrahymena cells (4–5 x 105 cells/ml) was grown in SPPA at 30°C overnight with shaking. Cells were harvested at 1,300 g for 5 min at 4°C and resuspended in a total volume of 15 ml of 10 mM Tris, pH 7.5. Protease inhibitors were added to a final concentration of: 0.7 mM PMSF, 3 μg/ml pepstatin, 3 μg/ml leupeptin, 30 μg/ml chymostatin, 10 μg/ml trans-epoxysuccinyl-L-leucylamido-(4guanidino)butane (E-64), and 15 μg/ml antipain. Cells were deciliated by adding 90 μl of 500 mM of dibucaine-HCl (Sigma Chemical Co.) followed by swirling. Just before all cells became nonmotile when viewed under a microscope (~2–3 min), 75 ml of ice-cold 10 mM Tris, pH 7.5, were added, and the suspension was spun at 1,900 g for 5 min at 4°C. The supernatant was spun again as above, and cilia were collected from the second supernatant by centrifugation at 14,000 g for 20 min at 4°C. Cilia were washed with 28 ml of ice-cold wash buffer (10 mM Hepes, pH 7.4, 100 mM NaCl, 4 mM MgCl2, and 0.1 mM EGTA) containing protease inhibitors, centrifuged at 14,000 g at 4°C for 10 min, resuspended (100 mg of wet ciliary pellet /ml) in cold wash buffer containing protease inhibitors, and frozen at -80°C. Ciliary protein concentrations were determined using the BCA assay kit (Pierce Chemical Co.).

    Protein Electrophoresis and Immunoblotting

    For -tubulin mutants, axonemal proteins extracted from cilia were mixed with an equal volume of 2x SDS sample buffer (Gaertig et al. 1995 ), boiled for 3–5 min, and run on 12% SDS–polyacrylamide gels (resolving gel at pH 8.8). Gels were blotted onto Immobilon-P PVDF membranes (Millipore Corp.) using a semi-dry transfer cell (Bio-Rad Laboratories). Blots were blocked in 3% BSA in TBS and probed with mouse primary mAbs. HRP rabbit anti–mouse IgG (1:2,000 dilution; Zymed Labs, Inc.) was used as a secondary antibody. Blots were developed with the Western blot chemiluminescence Reagent Plus kit (NEN Life Science Products). For ?-tubulin mutants, ciliary extracts (25–1,000 μg) were subjected to 10% SDS-PAGE, at pH 8.25 (Suprenant et al. 1985 ). Proteins were electrotransferred onto 0.45-μm nitrocellulose membranes at 20 V for 2 h. The blots were removed and blocked in PBST buffer (137 mM NaCl, 2.7 mM KCl, 4.3 mM Na2HPO4, 1.4 mM KH2PO4, and 0.1% Tween 20) containing 5% dried milk for 1 h at room temperature. Blots were incubated for 1 h at room temperature with the primary antibodies, membranes were washed in PBST and incubated with 1:3,000 dilution of goat anti–mouse or anti–rabbit antibodies linked to HRP (Amersham) for 1 h at room temperature. Blots were processed using the ECL system (Amersham), exposed to Fuji medical X-ray films, and the resulting bands were quantified by densitometry (ImageQuant Fast scan; Molecular Dynamics). For primary antibodies, we used the following mouse mAbs: TAP 952 directed against monoglycylated tubulin sites (1:1,000 dilution); AXO 49 directed against polyglycylated tubulins (1:667 dilution); general anti–-tubulin, DM1A (1:10,000 dilution; Amersham Pharmacia); general anti–?-tubulin, DM1B (1:10,000 dilution; Amersham Pharmacia); or polyclonal antibodies directed against 10xG peptide, anti-polyG at 1:100 dilution.

    Phenotypic Analyses

    Immunofluorescent detection of microtubules in Tetrahymena was performed as previously described (Gaertig et al. 1995 ). Sets of optical sections of individual cells were obtained using a laser scanning confocal microscope (model MRC 600; Bio-Rad Laboratories Inc.) and processed to obtain complete three-dimensional reconstruction. As a primary antibody, we used mAbs AXO 49, TAP 952, anti-polyG or anti-polyE antibodies generated against 10xE peptide at 1:100 dilution.

    To measure growth rates, cultures grown in 50 ml SPPA at 30°C with shaking (150 rpm) were counted at 4–5-h intervals using a Coulter counter (model ZF). To determine the rates of cell movement, cells were grown to density of 2 x 105 cells/ml. 10 μl of culture were placed on microscopic slides without coverslips and analyzed using an inverted microscope (model TE300; Nikon). Paths of moving cells were recorded for 4 s using a digital color CCD camera (model VI-470; Optronics) and IM-4000 software (version 3.46p; Image Analytics). The path lengths of moving cells was determined using the NIH Image program.

    Results

    -Tubulin Polyglycylation in T. thermophila Occurs at Multiple Sites and Is Not Essential

    E445 was originally shown to be a polyglycylation site in the -tubulin of Paramecium (Redeker et al. 1994 ) and it is conserved in Tetrahymena -tubulin (Fig 1). In ?-tubulin of Paramecium, polyglycylation can occur on several clustered Es downstream of the originally identified site, E437 (Vinh et al. 1999 ). Two Es also immediately follow E445 in Tetrahymena -tubulin, and we reasoned these Es could also be polyglycylated (Fig 1). To identify the sites, -tubulin genes mutated at these residues were used to rescue progeny of a mating between two -tubulin knockout heterokaryons (Hai and Gorovsky 1997 ). In the diploid, transcriptionally quiescent micronuclei of these knockout heterokaryons, the coding sequences of the ATU1 genes are replaced by the coding region of the neo gene, which confers resistance to paromomycin in Tetrahymena (Kahn et al. 1993 ). When the two heterokaryons conjugate, the parental macronuclei are destroyed, the new macronuclei of the progeny develop from the micronuclei and receive only the disrupted copies of the ATU1 gene. As a result, the progeny cells die when selected with paromomycin unless they are rescued with a functional ATU1 gene. Therefore, in the rescued progeny, the mutated version of the ATU1 gene is the only ATU1 gene that can be expressed. We constructed plasmids encoding mutated Tetrahymena ATU1 genes, in which the E445-447 codons were changed to aspartic acid (D) or alanine (A) codons. Mutant genes in which all three sites are substituted (DDD447 and AAA447) rescued ATU1 mating knockout heterokaryons.

    Figure 1. (A and B) COOH-terminal amino acid sequences of wild-type -tubulin (A), and wild-type ?-tubulin (B) of T. thermophila (Gaertig et al. 1993 ; McGrath et al. 1994 ) and P. tetraurelia (Dupuis 1992 ; Dupuis-Williams et al. 1996 ). Sites known to be polyglycylated in Paramecium are underlined. (C) COOH-terminal sequence of the mutant -tubulin of Tetrahymena in which the COOH-terminal end (positions 444–449) was replaced with the corresponding end of ?-tubulin (positions 436–443).

    To determine if the -tubulin in the transformants was polyglycylated, axonemal proteins were analyzed by immunoblotting with AXO 49, an mAb primarily recognizing polyglycine chains with three or more glycine residues (Gs; Bre et al. 1996 , Bre et al. 1998 ). A wild-type (WT) control derived from heterokaryons transformed by the WT gene displayed two bands of ~55 kD (Fig 2, middle). The top band comigrates with ?-tubulin, whereas the bottom band comigrates with -tubulin (Fig 2, bottom; Suprenant et al. 1985 ). The DDD447 and the AAA447 cells showed only one band corresponding to ?-tubulin (Fig 2, middle). The same result was obtained when immunoblots were stained with TAP 952, an mAb which primarily recognizes monoglycylated tubulin sites (Bre et al. 1996 , Bre et al. 1998 ; Fig 2, top). Thus, mutating E445-447 eliminated polyglycylation on -tubulin, indicating that this modification is not essential. Mutant ATU1 genes with single and double substitutions of E445-447 in all combinations also rescued mating heterokaryons but did not abolish -tubulin glycylation, indicating that each of the three Es serves as a modification site (data not shown).

    Figure 2. A Western blot analysis of axonemal proteins of T. thermophila WT and -tubulin mutant strains probed with mAbs TAP 952 (top) and AXO 49 (middle) specific for glycylated tubulin and a general anti–-tubulin mAb, DM1A (bottom).

    Mutant (AAA447 and DDD447) and WT cells are similar in appearance using light microscopy. No differences in growth rates were observed between the mutants and WT cells (Fig 6 A, left). The mutant cells were indistinguishable from WT cells in rates of swimming and cilia regeneration (data not shown). We used AXO 49 and TAP 952 mAbs to examine the cellular distribution of polyglycylated tubulins in WT and mutant (AAA447) cells. In interphase WT cells, AXO 49 mAbs reacted predominantly with somatic and oral cilia (Fig 3 B). TAP 952 antibodies reacted with somatic cilia, cortical microtubules, and intracytoplasmic microtubules during the entire cell cycle (Fig 3 D), as well as with micronuclear spindle microtubules during division (data not shown). Within cilia, TAP 952 reacted strongly with the ciliary tips (Fig 3 D). In contrast to AXO 49, the oral apparatus was not labeled by TAP 952 in nondividing cells, but oral cilia were labeled in newly formed oral primordia in cells approaching cell division (Fig 3 D). The overall labeling of AAA447 cells with both antibodies was indistinguishable from WT (Fig 3A and Fig C). The simplest explanation is that at the light microscopic level, the pattern of -tubulin glycylation is either the same as or is a subset of the pattern of ?-tubulin glycylation. Alternatively, -tubulin glycylation may still be restricted to certain organelles but may not be detectable by immunofluorescence because of its low level or lack of accessibility to modification-specific antibodies.

    Figure 3. Projections of AAA447 mutant (A and C) and WT cells (B and D) labeled with mAbs AXO 49 (A and B) and TAP 952 (C and D). Cells shown in A–C are in interphase, whereas the cell shown in D is in the stage before cell division. The oral apparatus was not labeled by TAP 952 in nondividing cells, but oral cilia were labeled in newly formed oral primordia in cells approaching cell division (D, WT cell). Similar observations were made for AAA447 cells (data not shown).

    Figure 4. (A) A Southern blot containing WT (lane 1) and btu2::bsr1 transformant genomic DNA (lane 2) digested with HindIII restriction endonuclease. The blot was probed with an - ATP–labeled fragment containing the 3'-untranslated region of BTU2. In WT, the probe detects a 3.0-kb fragment of the BTU2 gene. Gene disruption of BTU2 using the btu2::bsr1 fragment eliminates one HindIII site (H) and yields a fragment of 4.8 kb (see diagram on the right). The transformant DNA contains the 4.8-kb fragment and the endogenous 3-kb fragment, indicating that the transformant is heterozygous for btu2::bsr1/BTU2. (B) A Southern blot of total genomic DNA of WT (lane 1) and btu1::neo1 transformant (lane 2) probed with a 3'-untranslated fragment of BTU1. In WT, an 8-kb HindIII fragment was detected. In the transformed cell, a smaller 2.3-kb fragment was detected consistent with gene knockout of BTU1, which should introduce an additional HindIII site (H, see diagram on the right).

    Figure 5. (A and B) Western blots of WT and ?EDDD440 ciliary proteins probed with antiglycylated tubulin antibodies AXO 49 and TAP952. A series of dilutions of WT ciliary proteins were run that correspond to 0.25, 0.5, 1, 3, 5, and 10x the concentration of the ?EDDD440 mutant ciliary proteins analyzed on the same blot. Blots were processed using the ECL system and exposed to film. A scanning densitometer to quantify the amounts of antibody reactive epitopes. The shortest exposure for which the smallest amount of WT signal could be detected was used. A standard curve was prepared using the values for the WT extracts. For the blots shown, the relative intensities of mutant signals fell within the linear range of the standard curve. (C) A Western blot of WT and ?EDDD440 ciliary proteins probed with polyclonal antibodies against the surface antigen of T. thermophila, SerH3, and monoclonal anti–-tubulin antibodies DM1A. The same amounts of total ciliary proteins were loaded into each lane.

    Figure 6. (A, left) Growth curves of isogenic WT and DDD447 mutant cells. Cultures were grown at 30°C in SPPA medium with shaking. (A, right) Culture growth of isogenic WT and ?EDDD440 mutants. (B) Number of micronuclei (left) and macronuclei (right) per cell in ?EDDD440 mutant and isogenic WT cells.

    Creation of Knockout Heterokaryons for In Vivo Mutagenesis of ?-Tubulin

    To facilitate mutational analysis of ?-tubulin, we constructed heterokaryons for transformation rescue of the ?-tubulin genes. Two functionally redundant genes encode the major ?-tubulin of T. thermophila: BTU1 and BTU2 (Gaertig et al. 1993 ; Gu, L., J. Gaertig, and M.A. Gorovsky, unpublished results). We disrupted the BTU1 gene in the micronucleus by biolisitic bombardment using a BTU1::neo1 fragment (Fig 4 B). To disrupt the BTU2 gene, we used a btu2::bsr1 fragment in which BTU2 contains the blasticidin S resistance gene cassette (Fig 4 A). The two single knockout heterokaryons were used to construct two double knockout heterokaryons, DB2A and DB6B, having different mating types. When these two strains were crossed to each other and conjugation progeny were selected with bs (resistance conferred by the BTU2::bsr1 alleles in the new macronucleus), all exconjugants died, confirming that cells require either BTU1 or BTU2 for survival. However, bs-resistant progeny were obtained when the DB2A and DB6A mating cells were biolistically transformed with either a BTU1 or BTU2 fragment (Table 1 and results not shown). Thus, mating double knockout BTU heterokaryon strains can be used to introduce ?-tubulin genes allowing analysis of a series of mutations in residues of Tetrahymena ?-tubulin that are homologous to polyglycylation sites identified in Paramecium (see next section).

    Table 1. Sites, Transformation Efficiencies, and Gross Phenotypes of Wild Type and Various Mutations on ?-Tubulin

    Polyglycylation Sites on ?-Tubulin Are Essential for Survival

    E437 has been identified as a major site of polyglycylation on ?-tubulin in Paramecium (Redeker et al. 1994 ), and three additional downstream Es are also polyglycylated (Vinh et al. 1999 ). A series of mutations of BTU1 was made at the codons of the homologous sites at positions E437-E442 of Tetrahymena ?-tubulin (Table 1). We obtained viable, phenotypically normal progeny using fragments carrying D substitution at position 437 (?DEEE440) or at positions 437 and 438 (?DDEE440). No transformants were obtained for the ?DDDE440 and ?DDDD440 mutations (Table 1). To test whether E439 and E440 are essential, we prepared a fragment with the ?EEDD440 mutation and found that it transformed at high frequency, and that surviving cells appeared normal (Table 1). Thus, none of the four adjacent Es is essential by itself, but cells need to have at least two of the four consecutive Es for survival.

    Having determined that E440 alone is not sufficient, we tested whether any E in the region 437–439 was sufficient for survival. We repeatedly failed to obtain transformants using fragments carrying ?DEDD440 and ?DDED440 mutations. In contrast, when cells were transformed with the ?EDDD440 mutation, we obtained high frequency rescue. However, these cells were abnormal, showing defects in shape, movement, and growth (see next section). Thus, among the four adjacent Es, only E437 is sufficient for survival.

    In Paramecium ?-tubulin, an additional E (E441) located downstream of a stretch of three adjacent Es, was also found to be glycylated (Vinh et al. 1999 ). Alignment of the COOH-terminal peptides of Paramecium and Tetrahymena ?-tubulins suggests that E442 in Tetrahymena ?-tubulin is homologous to E441 of Paramecium (Fig 1). Therefore, E442 also might be polyglycylated in Tetrahymena and Tetrahymena cells could require both E437 and E442 for survival. To test this possibility, we used a mutant gene encoding the ?EDDDGD442 mutation to transform Tetrahymena (Table 1). Transformants carrying this mutation had an abnormal phenotype indistinguishable from that of ?EDDD440 mutants. Furthermore, mutants with a single substitution at position 442 (?EEEEGD442) were indistinguishable from WT. Thus, E437 is the only residue among the five COOH-terminal Es that is sufficient for survival.

    ?EDDD440 Mutants Have Reduced Mono- and Polyglycylation on ?-Tubulin and Increased Monoglycylation on -Tubulin

    To analyze the glycylation levels of mutant cells lacking likely polyglycylation sites, TAP 952 and AXO 49 mAbs were used to probe Western blots containing ciliary proteins. Initial analyses showed that the ?DEEE440, ?DDEE440, ?EEED440, ?EEDE440, ?EDDD440, and ?EEDD440 have polyglycylated epitopes on ?-tubulin and -tubulin (data not shown). Since the ?EDDD440 cells had a highly abnormal phenotype, the amount of glycylation in these cells was compared with that in WT cells. We first determined that mutant cilia contain similar amounts of ciliary proteins as WT cilia including -tubulin and a membrane protein, SerH3 (Fig 5 C). Then, Western blots containing a concentration series of WT or mutant ciliary proteins were probed with the glycylation-specific antibodies. Quantitation of Western blots showed that the amount of AXO 49 polyglycylated epitopes (three or more Gs) in the ?EDDD440 mutant ?-tubulin was only 24% (±6%, n = 4) of that in WT (Fig 5 A). Interestingly, the amount of AXO 49 epitopes were also somewhat reduced on the -tubulin of the ?EDDD440 mutant (53.5 ± 24% of WT, n = 4). When probed with TAP 952, which recognizes primarily monoglycylated sites, no signal was detectable on the ?-tubulin of ?EDDD440 mutants. In contrast, the TAP 952 epitopes increased more than twofold on -tubulin in the ?EDDD440 mutant (Fig 5 B).

    Growth, Motility, and Cytokinesis Are Affected in the ?EDDD440 Mutant

    During log phase, the generation time of ?EDDD440 cells was ~315 min compared with 250 min in WT cells (Fig 6 A, right). The average rate of movement of ?EDDD440 mutants is only 55% of WT cells, 14.2 μm/s (n = 92) compared with 25.7 μm/s (n = 94). No significant differences between length distributions of ciliary axonemes in WT and ?EDDD440 cells were detected (data not shown), indicating that polyglycylation did not affect cilia-based cell motility by affecting the extent of ciliary assembly.

    Confocal immunofluorescence analyses of WT and ?EDDD440 mutant cells were performed on cells stained with antitubulin antibodies (Fig 7). The size of ?EDDD440 mutant cells is more heterogeneous than WT cells (Fig 7 A); some are smaller (Fig 7 B) while others are bigger (Fig 7 D). Many ?EDDD440 cells also have an abnormally high number of nuclei (Fig 6 C and 7 D). In this mutant, we also observed cells in which nuclei had divided but apparently had failed to complete cytokinesis (Fig 7 C) as well as cells with a cleavage furrow that appeared improperly positioned and shifted toward one end of the cell (Fig 7E and Fig F). All of these phenotypes are consistent with defects in cytokinesis. The unusually small and large cells could be explained by unequal division, while large cells with multiple nuclei could result from cytokinesis failure.

    Figure 7. Confocal immunofluorescence images of wild type and ?EDDD440 mutants. (A–D) Cells were labeled with anti-polyG antibodies (green) and stained with propidium iodide (red). (E and F) The same cells were stained with anti-polyE antibodies (green) and DAPI (blue). (A) Wild-type cells showing normal cell shape, size, and numbers of nuclei; B–F are ?EDDD440 mutant cells.

    Lethal Mutations in the Polyglycylation Region of ?-Tubulin Can Be Rescued by Replacing the -Tubulin COOH Terminus by the ?-Tubulin COOH Terminus

    As described above, no transformants were obtained with the ?DDDE440, ?DDDD440, ?DEDD440, and ?DDED440 mutations, suggesting that polyglycylation on ?-tubulin, unlike polyglycylation on -tubulin, is essential. Another possible explanation is that the structure of these mutant tubulins is disrupted so that they fail to assemble into microtubules. Alternatively, polyglycylation of - and ?-tubulin could have overlapping functions, and elimination of this modification from -tubulin may be tolerated because there appears to be less polyglycylation of - than ?-tubulin (Fig 2). If polyglycylation of -tubulin is quantitatively insufficient to support the essential function of this PTM, it should be possible to rescue a lethal ?-tubulin mutation by creating a stronger polyglycylation region on -tubulin. However, if the defect is due to alteration in ?-tubulin function unrelated to the effect of the mutation on polyglycylation, increasing the polyglycylation of -tubulin should have no effect.

    To discriminate between these alternatives, we constructed a chimeric -tubulin gene whose COOH terminus was replaced with the corresponding region from WT ?-tubulin gene (Fig 1 C). We attempted to rescue the progeny of a mating between the double BTU knockout heterokaryons with this -COOH terminus-? fragment alone, with ?DDDE440 alone, or with a mixture of the two fragments. While we did not obtain rescues with either fragment, we did obtain rescues (two transformants per 2.5 μg of each fragment) when both fragments were used simultaneously. Using PCR and DNA sequencing, we confirmed that the rescued transformants had incorporated both genes. Although the rescue frequency was low, it was in the range of frequencies of cotransformation observed previously for a drug-resistant variant of BTU1 gene (Gaertig et al. 1994b ), making it unlikely that the infrequent double transformants had acquired additional spontaneous mutations at other loci. The ?DDDE 440 + -COOH terminus-? cells grew more slowly than WT. But, unlike ?EDDD440 mutants, the ?DDDE440 + -COOH terminus-? cells have normal shapes and numbers of nuclei, indicating that cytokinesis is not significantly disturbed (data not shown).

    In WT cells, ?-tubulin is more highly polyglycylated than . However, when ciliary proteins, which were isolated from ?DDDE440 + -COOH terminus cells, were analyzed on Western blots probed with AXO 49, the amount of polyglycylation on -tubulin was higher than that on ?-tubulin, inverting the pattern seen in WT cells (Fig 8). The fact that in the ?DDDE440 mutant ?-tubulin is still glycylated indicates that at least one of the remaining COOH-terminal Es (E440 and E441) is also polyglycylatable. To test whether E440 is a polyglycylation site, we transformed the BTU heterokaryons with the ?DDDD440 gene and the -COOH terminus-? chimera. Viable double transformants were obtained, but the mutant ?-tubulin was still polyglycylated (Fig 8 A), indicating that remaining E442 is also a polyglycylation site. To test E442, we introduced the ?DDDDGD442 mutation in the presence of -COOH terminus-? chimera. Viable double mutants were obtained at a frequency comparable to the frequency of transformation for ?DDDE440 and ?DDDD440 genes. No transformants were obtained with the ?DDDDGD442 gene alone (Table 1), and all transformants obtained in the presence of the -COOH terminus ? chimeric had incorporated both the mutant ?-tubulin gene and the -COOH terminus ? chimeric gene. On a Western blot, no monoglycylation or polyglycylation was detected in the region of ?-tubulin while the -? tubulin chimera was hyperpolyglycylated (Fig 8 B). The simplest explanation of these observations is that E442 was the single remaining site of polyglycylation in the ?DDDD440 mutant. Thus, all five COOH-terminal Es of ?-tubulin (E437-442) can be polyglycylated and the essential function of polyglycylation can be transferred from ?- to the -tubulin subunit. The phenotype of the ?DDDD440 and ?DDDDGD442 mutants was similar to the phenotype of the ?DDDE440 mutants (slow growth and motility but normal morphology, data not shown).

    Figure 8. (A) Immunoblot of total cell proteins of WT, ?DDDE440 (+ -COOH terminus-? chimera) and ?DDDD440 (+ -COOH terminus ?) chimera mutants probed with the polyclonal anti-polyG antibodies (top), monoclonal anti–?-tubulin antibodies, DM1B (middle), and monoclonal anti–-tubulin antibodies, DM1A (bottom). Approximately equal amounts of total cell proteins were loaded in each lane. (B) A Western blot of ciliary proteins of the ?DDDDGD442 (+ -COOH terminus-? chimera) and ?DDDE440 (+ -COOH terminus-? chimera) mutants probed with several antitubulin antibodies. Equal amounts of total proteins were loaded into each lane. The same blot was probed sequentially using the ECL system. After each exposure, the antibodies were removed, and the blot was reprobed according to manufacturer's instructions. The following antibodies were used: anti–?-tubulin, DM1B; anti–-tubulin, DM1A; antimonoglycylated tubulin, TAP 952; and antipolyglycylated tubulin, AXO 49.

    Discussion

    We examined the role of the - and ?-tubulin polyglycylation sites in Tetrahymena using in vivo mutagenesis and immunochemical analysis with tubulin glycylation–specific antibodies. The two mAbs used, AXO 49 and TAP 952, specifically recognize glycylated tubulin polypeptides (Callen et al. 1994 ; Bre et al. 1996 ). TAP 952 has high affinity towards tubulin peptides carrying a single G attached to one or several Es of the peptide backbone, whereas AXO 49 reacts with polyglycine chains consisting of 3 or more Gs (Bre et al. 1998 ). Therefore, the two antibodies should detect most polyglycylated tubulins with the possible exception of those modified with 2 G residues per site. We found that -tubulin polyglycylation detectable by these modification-specific antibodies can be eliminated completely by simultaneously replacing E445, E446, and E447 by aspartates or alanines. These observations argue that -tubulin polyglycylation is not essential.

    In the light of evolutionary conservation of tubulin glycylation in flagellated cells from evolutionarily divergent protists (Weber et al. 1996 ) to humans (Bre et al. 1996 ), it was surprising that the absence of -tubulin polyglycylation in Tetrahymena had no observable phenotype. However, the polyglycylation pattern of ?-tubulin appears to overlap with that of -tubulin, suggesting that polyglycylation of - and ?-tubulin is redundant and that polyglycylated ?-tubulin enabled the cells lacking -tubulin glycylation to function normally. Our genetic analyses argue strongly that this is the case. In contrast to -tubulin, the polyglycylation sites on ?-tubulin are essential for survival in Tetrahymena. We showed that in Tetrahymena, elimination of five, four, or three glutamic acids homologous to the polyglycylation sites previously identified in Paramecium (Vinh et al. 1999 ) was either lethal or caused slow growth, slow motility, and defects in cell division, associated with considerable reduction of ?-tubulin polyglycylation. The COOH-terminal Es of tubulins also can be polyglutamylated (see Introduction) in both axonemal and cytoplasmic microtubules, and this PTM was also detected in ciliates (Bre et al. 1994 ). Injection of an mAb specific for glutamylated tubulins led to a disassembly of centrioles in mammalian cells (Bobinnec et al. 1998 ). Thus, the dramatic phenotypic changes caused by substitutions of E437-E442 in ?-tubulin could conceivably result from either the reduction in the level of polyglycylation or polyglutamylation, or from a combination of both. However, in Paramecium, polyglutamylation must be a quantitatively minor modification since it was not detected by mass spectrometry, and only a very faint signal was detected on a Western blot with antiglutamylated tubulin antibodies (Bre et al. 1994 , Bre et al. 1998 ; Redeker et al. 1994 ). It is likely that the same holds true for Tetrahymena that belong to the same class of ciliates as Paramecium (Baroin-Tourancheau et al. 1992 ), and both ciliates have a similar design of their cytoskeletons (Fleury et al. 1992 ). A third possibility is that the phenotypes observed are induced by the change in the primary sequence of ?-tubulin. However, this explanation also seems unlikely because all ?-substitutions were charge conserving and all of the sites could be mutated singly without detectably affecting the phenotype. This argument is strengthened by the observation that lethal mutations of the polyglycylation region on ?-tubulin can be expressed when the COOH terminus of -tubulin is replaced by the COOH terminus of ?-tubulin in the same cells. Therefore, the ?-tubulin mutations that reduce or eliminate polyglycylation do not affect the ability of ?-tubulin to form tubulin dimers or to polymerize into microtubules.

    The polyglycylation-deficient mutants (?EDDD440) grow slowly, have reduced motility, and frequently arrest at cytokinesis. These phenotypes could reflect either a role for polyglycylation in MT assembly or a function in already assembled microtubules. Our data argue that polyglycylation is not required for the assembly of microtubules. First, the ?EDDD440 mutant assembles all classes of microtubular organelles found in vegetative cells. Furthermore, cilia that have highly glycylated microtubules in WT cells, have normal lengths in ?EDDD440 mutant cells. A postassembly function for polyglycylation is supported by an experiment in Drosophila where truncation of the COOH-terminal 15 amino acids of ?2 tubulin, the only isotype expressed in male germ cells, had no effect on either cytoplasmic or axonemal doublet MTs assembly in vivo (Fackenthal et al. 1993 ). However, the microtubules assembled from the truncated Drosophila ?-tubulin often failed to arrange properly within the axonemes (Fackenthal et al. 1993 ). These data suggest that the COOH terminus of ?-tubulin, and its modifications, are required for either the supramolecular organization of microtubules within organelles such as the axoneme or the function of already assembled microtubules.

    The dramatic reduction in the motility in the ?EDDD440 mutant is consistent with involvement of polyglycylation in regulation of ciliary beating in vivo. Because the length and number of mutant cilia are normal, the simplest explanation of the observation that ?EDDD440 mutant cells show a dramatic reduction in the rate of cell movement is that their cilia beat more slowly. Ciliary bending is based on limited sliding of adjacent peripheral doublets caused by ATP activation of dynein arms (Satir and Barkalow 1996 ). Polyglycylation could play a role in the regulation of ciliary dynein activity. Consistent with this, antipolyglycylation antibodies inhibited flagellar beating in reactivated permeabilized spermatozoa (Bre et al. 1996 ).

    The ?EDDD440 mutants also show defects in cytokinesis. Thus, polyglycylation may regulate the interaction of MTs with molecular motors involved in cytokinesis. Several kinesin motors were implicated in cytokinesis (Williams et al. 1995 ; Ohkura et al. 1997 ; Raich et al. 1998 ). Alternatively, polyglycylation of certain MTs may be more directly involved in determining cleavage furrow positioning and/or progression without any involvement of MAPs.

    An important aspect of our studies concerns the respective roles of - and ?-tubulin subunits in MTs. It is not understood why the functional subunit of MTs is a heterodimer. A heterodimer is not needed to establish polarity of a polymer since monomeric actin can form polarized filaments. Although both subunits bind a single GTP nucleotide, only GTP bound to ?-tubulin undergoes hydrolysis shortly after subunit addition to an MT end, and GDP–?-tubulin plays an important role in MT disassembly (Desai and Mitchison 1998 ). Studies in which MTs were decorated in vivo with purified kinesin heads showed that they bind mostly to ?-tubulin with some overlap with -tubulin (Song and Mandelkow 1993 ; Hirose et al. 1996 ; Downing and Nogales 1998 ). Thus, the ? subunit appears to be more important in both regulation of MT dynamics and interaction with motors. Our results are consistent with this model. We were able to eliminate detectable polyglycylation from the -subunit COOH terminus, but we could not obtain viable cells containing similarly mutated sites with ?-tubulin. Thus, the -subunit may be essential for the correct formation of a microtubule wall, but may play a less important role after microtubule assembly.

    Our results also suggest there are functional interactions between the COOH termini of the - and ?-tubulin subunits. In the ?EDDD440 mutant, reduction in the extent of mono- and polyglycylation on ?-tubulin was associated with a twofold increase in the extent of monoglycylation on -tubulin. Moreover, we showed that all sites of polyglycylation on ?-tubulin can be eliminated when the COOH-terminal end of -tubulin is altered to encode the corresponding sequence of ?-tubulin. Thus, an essential function of the COOH terminus of ?-tubulin can be transferred onto -tubulin. The cause of lethality of the polyglycylation region mutations on ?-tubulin appears to be an insufficient total level of polyglycylation, which can be compensated if similar sites are created on -tubulin. To our knowledge, this is the first demonstration of redundant function of the two subunits of the tubulin heterodimer.

    References

    Alexander, J.E., Hunt, D.F., Lee, M.K., Shabanowitz, J., Michel, H., Berlin, S.C., Macdonald, T.L., Sundberg, R.J., Rebhun, L.I., Frankfurter, A. 1991. Characterization of post-translational modifications in neuron-specific class III ?-tubulin by mass spectrometry. Proc. Natl. Acad. Sci. USA 88:4685-4689.

    Argarana, C.E., Barra, H.S., Caputto, R. 1978. Release of -tyrosine from tubulinyl--tyrosine by brain extract. Separation of a carboxypeptidase from tubulin tyrosine ligase. Mol. Cell. Biochem. 19:17-22.

    Baroin-Tourancheau, A., Delgado, P., Perasso, R., Adoutte, A. 1992. A broad molecular phylogeny of ciliates: identification of major evolutionary trends and radiations within the phylum. Proc. Natl. Acad. Sci. USA 89:9764-9768.

    Bobinnec, Y., Khodjakov, A., Mir, L.M., Rieder, C.L., Eddé, B., Bornens, M. 1998. Centriole disassembly in vivo and its effect on centrosome structure and function in vertebrate cells. J. Cell Biol. 143:1575-1589.

    Bré, M.H., de Nechaud, B., Wolff, A., Fleury, A. 1994. Glutamylated tubulin probed in ciliates with the monoclonal antibody GT335. Cell Motil. Cytoskelet 27:337-349.

    Bré, M.H., Redeker, V., Quibell, M., Darmanaden-Delome, J., Bressac, C., Cosson, J., Huitore, P., Schmitte, J.-M., Rossier, J., Johnson, T., Adoutte, A., Levilliers, N. 1996. Axonemal tubulin polyglycylation probed with two monoclonal antibodies: widespread evolutionary distribution, appearance during spermatozoan maturation and possible function in motility. J. Cell Sci. 109:727-738.

    Bré, M.H., Redeker, V., Vinh, J., Rossier, J., Levilliers, N. 1998. Tubulin polyglycylation: differential posttranslational modification of dynamic cytoplasmic and stable axonemal microtubules in Paramecium. Mol. Biol. Cell 9:2655-2665.

    Callen, A.-M., Adoutte, A., Andrew, J.M., Baroin-Tourancheau, A., Bré, M.-H., Ruiz, P.C., Clérot, J.-C., Delgado, P., Fleury, A., Jeanmaire-Wolf, R., Viklicky, V., Villalobo, E., Levilliers, N. 1994. Isolation and characterization of libraries of monoclonal antibodies directed against various forms of tubulin in Paramecium. Biol. Cell 81:95-119.

    Caron, J.M. 1997. Posttranslational modification of tubulin by palmitoylation. I. In vivo and cell-free studies. Mol. Biol. Cell 8:621-636.

    Cassidy-Hanley, D., Bowen, J., Lee, J., Cole, E.S., VerPlank, L.A., Gaertig, J., Gorovsky, M.A., Bruns, P.J. 1997. Germline and somatic transformation of mating Tetrahymena thermophila by particle bombardment. Genetics 146:135-147.

    Conzelmann, K.K., Helftenbein, E. 1987. Nucleotide sequence and expression of two beta-tubulin genes in Stylonychia lemnae. J. Mol. Biol. 198:643-653.

    Desai, A., Mitchison, T.J. 1998. Tubulin and FtsZ structures: functional and therapeutic implications. Bioessays 20:523-527.

    Downing, K., Nogales, E. 1998. Tubulin and microtubule structure. Curr. Opin. Cell Biol 10:16-22.

    Dupuis, P. 1992. The ?-tubulin genes of Paramecium are interrupted by two 27 bp introns. EMBO (Eur. Mol. Biol. Organ.) J. 11:3713-3719.

    Dupuis-Williams, P., Klotz, C., Mazarguil, H., Beisson, J. 1996. The tubulin gene family of Paramecium: characterization and expression of the alpha PT1 and alpha PT2 genes which code for alpha-tubulins with unusual C-terminal amino acids, GLY and ALA. Biol. Cell 87:83-93.

    Eddé, B., Rossier, J., Le Caer, J.-P., Desbruyeres, E., Gros, F., Denoulet, P. 1990. Posttranslational glutamylation of -tubulin. Science 247:83-85.

    Eipper, B.A. 1972. Rat brain microtubule protein: purification and determination of covalently bound phosphate and carbohydrate. Proc. Natl Acad. Sci. USA 69:2283-2287.

    Fackenthal, J.D., Turner, F.R., Raff, E.C. 1993. Tissue-specific microtubule functions in Drosophila spermatogenesis require the ?2 isotype-specific carboxy terminus. Dev. Biol. 158:213-227.

    Fleury, A., Delgado, P., Iftode, F., Adoutte, A. 1992. Molecular phylogeny of ciliates: what does it tell us about the evolution of the cytoskeleton and of developmental strategies? Dev. Genet. 13:247-254.

    Gaertig, J., Thatcher, T.H., McGrath, K.E., Callahan, R.C., Gorovsky, M.A. 1993. Perspectives on tubulin isotype function and evolution based on the observations that Tetrahymena thermophila microtubules contain a single - and ?-tubulin. Cell Motil. Cytoskelet. 25:243-253.

    Gaertig, J., Gu, L., Hai, B., Gorovsky, M.A. 1994a. High frequency vector-mediated transformation and gene replacement in Tetrahymena. Nucleic Acids Res. 22:5391-5398.

    Gaertig, J., Thatcher, T.H., Gu, L., Gorovsky, M.A. 1994b. Electroporation-mediated replacement of a positively and negatively selectable ?-tubulin gene in Tetrahymena thermophila. Proc. Natl. Acad. Sci. USA 91:4549-4553.

    Gaertig, J., Cruz, M.A., Bowen, J., Gu, L., Pennock, D.G., Gorovsky, M.A. 1995. Acetylation of lysine 40 in -tubulin is not essential in Tetrahymena thermophila. J. Cell Biol. 129:1301-1310.

    Hai, B., Gorovsky, M.A. 1997. Germ-line knockout heterokaryons of an essential alpha-tubulin gene enable high-frequency gene replacement and a test of gene transfer from somatic to germ-line nuclei in Tetrahymena thermophila. Proc. Natl. Acad. Sci. USA. 94:1310-1315.

    Hirose, K., Lockhart, A., Cross, R.A., Amos, L. 1996. Three-dimensional cryoelectron microscopy of dimeric kinesins and ncd motor domains on microtubules. Proc. Natl. Acad. Sci. USA 93:9539-9544.

    Kahn, R.W., Andersen, B.H., Brunk, C.F. 1993. Transformation of Tetrahymena thermophila by microinjection of a foreign gene. Proc. Natl. Acad. Sci. USA 90:9295-9299.

    Kunkel, T.A. 1985. Rapid and efficient site-specific mutagenesis without phenotypic selection. Proc. Natl. Acad. Sci. USA 82:488-492.

    L'Hernault, S.W., Rosenbaum, J.L. 1985. Chlamydomonas -tubulin is posttranslationally modified by acetylation on the E-amino group of a lysine. Biochemistry. 24:473-478.

    Levilliers, N., Fleury, A., Hill, A.M. 1995. Monoclonal and polyclonal antibodies detect a new type of post-translational modification of axonemal tubulin. J. Cell Sci. 108:3013-3028.

    Luduena, R.F. 1998. Multiple forms of tubulin: different gene products and covalent modifications. Int. Rev. Cytol. 178:207-274.

    MacRae, T.H. 1997. Tubulin post-translational modifications enzymes and their mechanisms of action. Eur. J. Biochem. 244:265-278.

    Mary, J., Redeker, V., Le Caer, J.-P., Promé, J.-C., Rossier, J. 1994. Class I and IVa ?-tubulin isotypes expressed in adult mouse brain are glutamylated. FEBS (Fed. Eur. Biochem. Soc.) Lett. 353:89-94.

    Mary, J., Redeker, V., Le Caer, J.-P., Rossier, J., Schmitter, J.M. 1996. Posttranslational modifications in the C-terminal tail of axonemal tubulin from sea urchin sperm. J. Biol. Chem. 271:9928-9933.

    McGrath, K.E., Yu, S.M., Heruth, D.P., Kelly, A.A., Gorovsky, M.A. 1994. Regulation and evolution of the single alpha-tubulin gene of the ciliate Tetrahymena thermophila. Cell Motil. Cytoskelet 27:272-283.

    Multigner, L., Pignot-Paintrand, I., Saoudi, Y., Job, D., Plessmann, U., Rüdiger, M., Weber, K. 1996. The A and B tubules of the outer doublets of sea urchin sperm axonemes are composed of different tubulin variants. Biochemistry 35:10862-10871.

    Ohkura, H., Torok, T., Tick, G., Hoheisel, J., Kiss, I., Glover, D.M. 1997. Mutation of a gene for a Drosophila kinesin-like protein, KLP38B, leads to a failure of cytokinesis. J. Cell Sci. 110:945-954.

    Orias, E., Bruns, P.J. 1976. Induction and isolation of mutants in Tetrahymena. In Prescott D.M., ed. Methods in Cell Biology. Vol. 13. New York, Academic Press, 247-282.

    Raich, W.B., Moran, A.N., Rothman, J.H., Hardin, J. 1998. Cytokinesis and midzone microtubule organization in Caenorhabditis elegans require the kinesin-like protein ZEN-4. Mol. Biol. Cell 9:2037-2049.

    Redeker, V., Melki, R., Promé, D., Le Caer, J.-P., Rossier, J. 1992. Structure of tubulin C-terminal domain obtained by subtilisin treatment. The major - and ?-tubulin isotypes from pig brain are glutamylated. FEBS (Fed. Eur. Biochem. Soc.) Lett. 313:185-192.

    Redeker, V., Levilliers, N., Schmitter, J.-M., Le Caer, J.-P., Rossier, J., Adoutte, A., Bré, M.-H. 1994. Polyglycylation of tubulin: a post-translational modification in axonemal microtubules. Science 266:1688-1691.

    Rüdiger, M., Plessman, U., Kl?ppel, K.-D., Wehland, J., Weber, K. 1992. Class II tubulin, the major brain ? tubulin isotype is polyglutamylated on glutamic acid residue 435. FEBS (Fed. Eur. Biochem. Soc.) Lett. 308:101-105.

    Rüdiger, M., Plessman, U., Rüdiger, A.-H., Weber, K. 1995. ?-Tubulin of bull sperm is polyglycylated. FEBS (Fed. Eur. Biochem. Soc.) Lett. 264:147-151.

    Satir, P., Barkalow, K.L. 1996. Cilia: structure and molecular biology. In Hausmann K., Bradbury P.C., eds. Ciliates: Cells as Organism. Stuttgart, Gustav Fischer, 355-377.

    Silflow, C.D. 1991. Why do tubulin gene families lack diversity in flagellate/ciliate protists? Protoplasma 164:9-11.

    Song, Y.-H., Mandelkow, E. 1993. Recombinant kinesin motor domain binds to ?-tubulin and decorates microtubules with a B surface lattice. Proc. Natl. Acad. Sci. USA 90:1671-1675.

    Suprenant, K.A., Hays, E., LeCluyse, E., Dentler, W.L. 1985. Multiple forms of tubulin in the cilia and cytoplasm of Tetrahymena thermophila. Proc. Natl. Acad. Sci. USA 82:6908-6912.

    Sutoh, K. 1993. A transformation vector for Dictyostelium discoideum with a new selectable marker bsr. Plasmid 30:150-154.

    Vinh, J., Langridge, J.I., Bré, M.-H., Levilliers, N., Redeker, V., Loyaux, D., Rossier, J. 1999. Structural characterization by tandem spectroscopy of the posttranslational modifications of tubulin. Biochemistry 38:3133-3139.

    Weber, K., Schneider, A., Muller, N., Plessman, U. 1996. Polyglycylation of tubulin in the diplomonad Giardia lamblia, one of the oldest eukaryotes. FEBS (Fed. Eur. Biochem. Soc.) Lett. 393:27-30.

    Williams, B.C., Riedy, M.F., Williams, E.V., Gatti, M., Goldberg, M.L. 1995. The Drosophila kinesin-like protein KLP3A is a midbody component required for central spindle assembly and initiation of cytokinesis. J. Cell Biol. 129:709-723.(Lu Xiaa, Bing Haib, Yan Gaoa, Dylan Burn)