Download Saidi_etal_FINAL_220911

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

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

Document related concepts

Cell wall wikipedia , lookup

Auxin wikipedia , lookup

Evolutionary history of plants wikipedia , lookup

Plant nutrition wikipedia , lookup

Botany wikipedia , lookup

Plant use of endophytic fungi in defense wikipedia , lookup

History of botany wikipedia , lookup

Plant defense against herbivory wikipedia , lookup

Meristem wikipedia , lookup

Plant reproduction wikipedia , lookup

Plant secondary metabolism wikipedia , lookup

Plant morphology wikipedia , lookup

Arabidopsis thaliana wikipedia , lookup

Plant breeding wikipedia , lookup

Plant physiology wikipedia , lookup

Plant stress measurement wikipedia , lookup

Plant ecology wikipedia , lookup

Glossary of plant morphology wikipedia , lookup

Sustainable landscaping wikipedia , lookup

Perovskia atriplicifolia wikipedia , lookup

Plant evolutionary developmental biology wikipedia , lookup

Transcript
1
Function and evolution of 'green' GSK3/Shaggy-like kinases
2
3
Younousse Saidi, Timothy J. Hearn and Juliet C. Coates
4
School of Biosciences, University of Birmingham, Edgbaston, Birmingham, B15 2TT,
5
United Kingdom
6
Corresponding author: Coates, J.C. ([email protected]).
7
8
Glycogen synthase kinase 3 (GSK3) proteins, also known as SHAGGY-like kinases,
9
have many important cell signalling roles in animals, fungi and amoebae. In
10
particular, GSK3s participate in key developmental signalling pathways and also
11
regulate the cytoskeleton. GSK3-encoding genes are also present in all land plants
12
and in algae and protists, raising questions about possible ancestral functions in
13
eukaryotes. Recent studies have revealed that plant GSK3 proteins are actively
14
implicated in hormonal signalling networks during development as well as in biotic
15
and abiotic stress responses. In this review, we outline the mechanisms of
16
Arabidopsis GSK3 action, summarize GSK3 functions in dicot and monocot
17
flowering plants, and speculate on the possible functions of GSK3s in the earliest-
18
evolving land plants.
19
20
Multifunctional proteins with conserved structure
21
Glycogen synthase kinase 3 (GSK3) was originally characterized in the animal insulin
22
signalling pathway as a serine/threonine kinase that phosphorylates glycogen synthase,
23
the enzyme responsible for the final step in the synthesis of glycogen [1]. GSK3
24
functions in many developmental processes, including cell fate specification,
1
25
cytoskeleton movements and programmed cell death (reviewed in [2]), with roles in
26
human diseases, including cancer and Alzheimer’s (reviewed in [3]). In particular, GSK3
27
is a central player in the animal Wnt signalling pathway, where extracellular Wnt signals
28
lead to the inactivation of GSK3. This blocks the activity of GSK3 towards β-catenin (see
29
Glossary), which as a result is no longer degraded by the proteasome, but can build up in
30
the nucleus and regulate target gene expression [4]. Thus, GSK3 is crucial for
31
developmental patterning [5], a role that appears to be conserved in Dictyostelium
32
discoideum [6]. Mammalian GSK3 exists as two isoforms, encoded by separate genes,
33
GSK3α and GSK3β, which have a conserved kinase domain but divergent N- and C-
34
terminal domains, which are important for regulation of function [7].
35
GSK3 substrates are diverse and most of them require phosphorylation by another
36
kinase before being phosphorylated by GSK3. This is termed as a 'priming'
37
phosphorylation, which positions the substrate in a suitable configuration for
38
phosphorylation by GSK3. A phosphate-binding pocket in GSK3 interacts with the
39
primed phosphorylation [8]. In GSK3β, the phosphate-binding pocket is defined by
40
arginine 96, arginine 180 and lysine 205 [8] (Figure 1). GSK3 is an ancient kinase, with
41
homologues found in all eukaryotes studied to date. Unlike in animals and Dictyostelium,
42
land plant GSK3s are encoded by relatively large multigene families whose members
43
share high sequence similarity. In all angiosperm GSK3s so far analysed, the phosphate-
44
binding pocket residues are identical to those in GSK3, suggesting that plant GSK3s can
45
phosphorylate primed substrates [9] (Figure 1). Accordingly, a proteomic study identified
46
ten Arabidopsis thaliana (Arabidopsis) (proteins with putative GSK3 phosphorylation
47
sites: five of these were phosphorylated at the corresponding priming site [10].
2
48
In Arabidopsis there are ten GSK3 homologues, also termed Shaggy Kinases
49
(AtSK or ASK) in reference to the Drosophila GSK3 homologue [11]. Nomenclature of
50
Arabidopsis GSK3s can be confusing. A full complement of names for each Arabidopsis
51
protein, along with names for other plant GSK3s that have been studied, is given in Table
52
1. In the last decade, substantial progress has been made in understanding how plant
53
GSK3s perform their diverse functions. In the following sections we will describe
54
currently known plant GSK3 functions, discuss the molecular mechanisms of GSK3
55
action in plants, and highlight possible GSK3 functions in early-evolving land plants as
56
an exciting area for future research developments.
57
58
GSK3s and brassinosteroid signalling in flowering plants
59
Steroid hormone signalling is found in plants, animals and fungi and, therefore, might be
60
predicted to have an ancient evolutionary origin [12]. However, the proteins involved in
61
plant steroid signalling belong to plant-specific protein families, thus indicating a
62
separate origin for steroid signalling in the plant lineage. Brassinosteroid (BR) signalling
63
modulates a range of physiological responses in flowering plants, including cell
64
expansion, greening, flowering time, fertility and differentiation of vascular tissue
65
(reviewed in [13-16]).
66
In Arabidopsis, BRs are perceived by a plasma membrane-localized receptor
67
kinase, BRI1 [13-15]. The downstream signal transduction involves a gene isolated by
68
forward genetic studies, BRASSINOSTEROID-INSENSITIVE 2 (BIN2). Dominant
69
mutations in bin2 result in brassinosteroid-insensitive dwarf plants. BIN2 encodes a
70
GSK3 with a catalytic domain sharing ~70% identity with animal GSK3β [17-20] (Figure
3
71
1). Gain-of-function bin2 point mutations and over-expression of BIN2 all cause the
72
same BR insensitivity phenotype, confirming BIN2 as an inhibitor of BR signalling [17-
73
20].
74
In the absence of a BR signal, active BIN2 negatively regulates BR-specific
75
transcription factors, BRASSINAZOLE RESISTANT 1 (BZR1) and bri1-EMS-
76
SUPPRESSOR 1 (BES1/BZR2) (Figure 2). BIN2 phosphorylates BZR1/BES1 on many
77
serine and threonine residues, which leads to their proteasomal degradation [21-23],
78
reminiscent of the role of animal GSK3 during both Wnt- and Hedgehog signalling [4,
79
24]. BIN2 can also promote BZR1/BES1 nuclear export, thus inhibiting BZR1/BES1
80
binding to DNA [25-26]. Again, this is a role also demonstrated by activated
81
(phosphorylated (Figure 1)) animal and amoebal GSK3, which control the nuclear export
82
of the transcription factors NF-ATc and StatA, respectively [27-28]. Interaction between
83
BIN2 and BZR1/BES1 is via a specific docking mechanism not seen in other GSK3
84
substrates [29], and independent of prime phosphorylation or a scaffold protein [21]. In
85
the presence of BR, BIN2 activity towards BES1/BZR1 is inhibited and, consequently,
86
unphosphorylated BZR1 and BES1 are no longer degraded and can direct transcription of
87
BR-responsive genes (Figure 2) [13-15].
88
The exact mechanism by which the BR signal is transduced from the plasma
89
membrane to BIN2 has now been elucidated [13-15,30-32] and it demonstrates novel
90
modes of inhibition of GSK3 activity. BR is perceived by the BRI1 brassinosteroid
91
receptor and BAK1 (BRI1-associated receptor kinase) co-receptor (Figure 2). BR binding
92
to BRI1 enables transphosphorylation of BRI1 and BAK1, and activation of BRI1. BRI1
93
then phosphorylates BR-signalling kinases (BSKs) at the plasma membrane.
4
94
Phosphorylated BSK binds and activates BSU1 phosphatases, which dephosphorylate
95
tyrosine 200 (Y200) in BIN2, inactivating its kinase activity [31] (Figure 2).
96
Interestingly, the dominant bin2-1 (E263K) mutation (Figure 1) renders BIN2
97
hyperactive by blocking its dephosphorylation by BSU1 [31]. In response to BR, BIN2 is
98
also degraded by the proteasome, whereas the bin2-1 mutant protein is considerably
99
stabilized [30]. Whether it is the Y200-dephosphorylated form of BIN2 that is
100
specifically targeted for degradation has yet to be determined. Y200 in BIN2 is
101
homologous to Y216 in human GSK3β (Figure 1), a tyrosine residue absolutely required
102
for GSK3 activity. The homologous tyrosine residue in other AtGSK3s is also
103
phosphorylated in vivo, and this phosphorylation is required for kinase activity [10]. In
104
GSK3β, this residue is autophosphorylated by GSK3 itself just after the protein is
105
synthesized, with the help of the Hsp90 chaperone protein [33]. BSU1-mediated
106
dephosphorylation of Y200 in BIN2 is the first example of direct dephosphorylation of
107
the conserved tyrosine residue of GSK3 as a means of negatively regulating kinase
108
activity. Furthermore, most plant GSK3s (all except clade III kinases, Table 1) do not
109
contain the N-terminal serine 9 residue found in GSK3β that is key to regulating GSK3β
110
inhibition (Figure 1) (reviewed in [2-3]). It is, therefore, possible that a different
111
mechanism exists in plants for GSK3 inactivation.
112
The mechanism for brassinosteroid perception and transduction involves GSK3
113
homologues functioning redundantly for the negative regulation of BR responses [26,34-
114
35]. Experiments using loss- and gain-of-function mutants showed that two other GSK3
115
homologues, AtSK22/BIL1 and AtSK23/BIL2 (Table 1), function in the same way as
116
BIN2 during BR signalling [35]. Together these genes make up sub-clade II of the
5
117
Arabidopsis GSK3s in the flowering plant phylogeny [9,36] (Table 1). Although the
118
triple mutant lines of BIN2, BIL1 and BIL2 show a BR-enhanced phenotype, analysis of
119
the levels of phosphorylated BZR1 showed that despite knockouts of all clade II GSK3s
120
there was still substantial phosphorylation of BZR1 [35]. This indicated a role for
121
additional GSK3(s) in the BR signalling pathway, which have been identified as
122
AtSK31/ASKθ [37], a member of sub-clade III, and possibly all three of the subclade I
123
GSK3s [31]. Therefore, at least seven out of the ten Arabidopsis GSK3 proteins are
124
implicated in BR signalling. Importantly, a chemical activator of BR signalling, bikinin,
125
specifically inhibits the same subset of seven Arabidopsis GSK3 proteins [38]. BIN2
126
orthologues from cotton (Gossypium hirsutum) and rice (Oryza sativa) have been isolated
127
and their overexpression in Arabidopsis has been shown to cause severe growth defects
128
similar to bin2 gain-of-function (BR-insensitive) mutants [39-40]. This suggests that
129
BIN2 orthologues encode isoforms that share a conserved function in BR signalling.
130
131
BIN2 at the crossroads between brassinosteroids, abscisic acid and auxin
132
One exciting finding is that GSK3 may mediate the crosstalk between BR signalling and
133
other hormone signalling pathways [41-42]. Abscisic acid (ABA) inhibits BR signalling,
134
enhances BES1 phosphorylation and induces the expression of BR-suppressed genes
135
[41]. This effect was alleviated when BIN2 activity was blocked using lithium chloride
136
(LiCl), a GSK3 inhibitor. The effect of ABA on BR signalling was independent of the
137
early BR perception events [41]. However, whether ABA has a direct effect on BIN2
138
phosphorylation or whether it affects the upstream BSK/BSU components remains to be
139
elucidated (Figure 2). Auxin, unlike ABA, has no effect on the phosphorylation state of
6
140
BES1 [43]. However, ARF2, an auxin response factor (ARF) that acts as a transcriptional
141
repressor, has numerous putative GSK3 recognition sites and the interaction between
142
BIN2 and ARF2 has been confirmed in yeast [42]. In vitro studies also showed that BIN2
143
is capable of phosphorylating ARF2 and consequently inhibiting its DNA-binding and
144
repressor activity [42]. These findings provide a molecular explanation for the effect of
145
BRs on the auxin response by suggesting that BIN2 facilitates binding of activator ARF
146
proteins by removing repressor ARFs from regulatory elements in the promoters of
147
auxin-responsive genes ([42] and Figure 2). It is possible that hormone signalling may
148
impinge upon plant GSK3s in additional ways. For example, AtSK31 gene expression is
149
upregulated by auxin in roots [44], although AtSK31 does not have a demonstrated role
150
in BR or auxin signalling.
151
152
GSK3 function in floral organs and cell expansion
153
Molecular and genetic analysis has confirmed roles for Arabidopsis GSK3s in the
154
development of flowers and reproductive organs. AtSK11 and AtSK12 show specific
155
strong expression in early floral meristems, which later becomes restricted to sepal
156
primordia, petals, carpels and the pollen-containing regions of the anthers [45]. Within
157
the carpels, AtSK11 and AtSK12 expression concentrates on adaxial sides (nearest the
158
central axis) of the carpels and in ovule primordia, particularly in the integuments and
159
megaspore mother cell [45]. A quantitative RT-PCR study of all ten Arabidopsis GSK3s
160
revealed that a third GSK3, AtSK31, was also expressed relatively highly in floral organs
161
[46]. Moreover, the AtSK31 protein localizes to the nuclei of developing tissues,
162
particularly in floral organs, gametophytes and embryos [47].
7
163
Consistent with the expression data, antisense reduction in either AtSK11 or
164
AtSK12 transcript levels results in disrupted cell division in the floral meristem, in
165
flowers with increased numbers of sepals and petals, and in abnormal carpel development
166
[45]. Interestingly, overexpression of different AtSK32 isoforms in wild-type Arabidopsis
167
causes varying effects [48]. Overexpression of either wild-type AtSK32 or a kinase-dead
168
mutant (K167A) resulted in no detectable phenotype [48]. However, overexpression of a
169
mutant specifically unable to phosphorylate primed GSK3 substrates (R178A; Figure 1)
170
resulted in plants with short hypocotyls and reduced cell elongation in floral organs.
171
Accordingly, AtSK32 (K167A)-overexpressing plants showed a marked downregulation
172
of the genes encoding several cell wall-modifying enzymes involved in cell elongation
173
[48]. This suggests that GSK3-mediated cell elongation responses may require the action
174
of primed GSK3 substrates. It is likely that in addition to floral cell expansion, GSK3s
175
may have a more general role in cell elongation in the plant given that BRs are
176
instrumental in controlling cell elongation and dominant bin2/ucu1 mutants show cell
177
elongation defects in leaves [19]. Given that GSK3 targets BZR1 and BES1 do not
178
require a priming phosphorylation [21,30], it seems that new targets of AtSK32 await
179
discovery.
180
181
GSK3 roles in stress responses
182
Work in several plant species has implicated GSK3s in abiotic stress responses, although
183
sometimes in conflicting ways [40,46,49-52] (Table 1). An up-regulation of GSK3
184
transcripts by salt stress has been reported in Arabidopsis (AtSK13, AtSK31 and AtSK42)
185
[46], wheat (Triticum aestivum) (TaGSK1) [53], rice (OsGSK1) [40] and sugarcane
8
186
(Saccharum officinarum) (SuSK) [52]. The above mentioned Arabidopsis GSKs are also
187
induced by osmotic stress whereas rice OsGSK1 transcripts show a strong reduction
188
following drought treatment [40]. Confirming the role of GSK3 as an active component
189
of the salt stress signalling pathway, Arabidopsis plants overexpressing AtGSK1
190
(AtSK22) show a significant upregulation of several salt-stress-responsive genes, even in
191
the absence of high salinity [49]. As a result, they exhibit an enhanced resistance to high
192
salt conditions. In another study, a rice mutant with a T-DNA insertion in OsGSK1
193
exhibited elevated transcript levels of specific stress-responsive genes and a strong
194
tolerance to high salt and drought treatments [40]. In Medicago sativa (alfalfa), the kinase
195
activity of a plastid localized GSK3 (MsK4) was rapidly increased under hyperosmotic
196
conditions [51]. The overexpression of MsK4 in Arabidopsis improved salt tolerance
197
through changes in carbohydrate metabolism [51]. MsK4 is most similar to AtSK41,
198
which does not respond to stress treatments [46]. However, AtSK31 shows a unique
199
upregulation in plants that have been subjected to a dark period [46], making it a
200
promising candidate for future studies.
201
GSK3s are also important in the plant response to biotic stress. M. sativa MsK1
202
shows a change in activity in response to plant defence elicitors. Interestingly, exposure
203
to cellulase rapidly inhibited MsK1 activity and triggered its proteasome-dependent
204
degradation [54]. The role of GSK3 in modulating the disease response was demonstrated
205
when over-expression of MsK1 in Arabidopsis reduced the pathogen-mediated activation
206
of AtMPK3 and AtMPK6 MAP kinases and exacerbated the susceptibility of the plant to
207
Pseudomonas syringae [54]. This suggests that MsK1 acts as negative regulator of the
208
basal defence response and that its elicitor-triggered regulation is essential to activate
9
209
downstream components of plant defence pathways [54]. Another M. sativa GSK3, WIG
210
(wound-induced gene), has a kinase activity that is transiently triggered following leaf
211
injury [55]. Although none of the Arabidopsis GSKs seem to be transcriptionally induced
212
by wounding [46], their kinase activity under mechanical injury has never been tested to
213
date. Overall, GSK3 family members seem to have distinct regulatory functions and their
214
activity appears to be modulated by various cues. How different stimuli modulate GSK3
215
activity during stress responses is largely unknown. One attractive hypothesis is that this
216
occurs through phosphorylation/dephosphorylation by upstream components but this has
217
yet to be demonstrated. Future research should also focus on isolating the targets of stress
218
response-related GSK3s, which should open new possibilities in the quest for crop
219
improvement.
220
The stress-responsive nature of GSK3s is evolutionarily ancient. GSK3 genes in
221
the yeast Saccharomyces cerevisiae, particularly MCK1, are involved in responses to
222
environmental stresses, including heat, cold, salt/osmotic stress, nutrient stress, oxidative
223
stress and metal ion stress [56-58]. The temperature-sensitive phenotype of a yeast
224
quadruple gsk3 mutant is rescued by expression of mammalian GSK3β [59], suggesting
225
further cross-kingdom conservation of functionality. Similarly, Arabidopsis AtSK22 can
226
rescue the salt-sensitive phenotype of the yeast mck1 mutant [60]. To thrive following
227
transition from water to land, plants had to acquire new stress-tolerant adaptations,
228
particularly in response to desiccation, light and temperature changes. For this reason, in
229
the next section we consider the GSK3s of algae and of plants that evolved early to life
230
on land.
231
10
232
Evolution of plant GSK3 function
233
GSK3 genes may have played an instrumental role in the acquisition of stress tolerance
234
mechanisms during the adaptation to life outside water because land plants are unusual in
235
possessing large GSK3 gene families. GSK3 gene diversification has occurred more than
236
once during land plant evolution [36,50]. In seed plants, three GSK3 clades are
237
represented in gymnosperms, with divergence into four clades occurring before (or early
238
in) angiosperm evolution [36,50]. The moss Physcomitrella patens diverged early from
239
the seed plant lineage around 470 million years ago, shortly after plants appeared on land.
240
The five Physcomitrella GSK3 genes discovered by cDNA library screening [50] fit into
241
a separate subclade that is most similar to flowering plant clades I and IV [36,50].
242
Physcomitrella GSK3s share 95% sequence identity across the whole protein [50],
243
suggesting a large degree of redundancy and possibly recent gene duplications in moss.
244
The complete sequencing of the Physcomitrella [61] and the lycophyte Selaginella
245
moellendorffii [62] genomes has made it possible to assess the full complement of GSK3
246
gene sequences in the early land plants. We can detect seven putative GSK3 homologues
247
in Physcomitrella, and two in Selaginella (J.C. Coates, unpublished). Our tentative
248
phylogenies suggest that the Selaginella GSK3 genes duplicated after the divergence of
249
Selaginella and Physcomitrella.
250
The functions of GSK3s in Physcomitrella and Selaginella are unknown.
251
However, as in Arabidopsis and rice, two Physcomitrella genes show increased transcript
252
abundance when exposed to sorbitol and polyethylene glycol [50] (Table 1), suggesting a
253
conserved role in osmotic stress responses. Interestingly, sequenced genomes of green
254
and red algae (Chlamydomonas reinhardtii, Volvox carteri, Porphyra umbilicalis and
11
255
Cyanidioschyzon merolae) each contain only one GSK3 gene, further suggesting that all
256
the 'green' GSK3 gene duplications occurred after the evolution of land plants.
257
Within the heterokont lineage, we can detect a single GSK3 homologue in the
258
sequenced genomes of each of: Ectocarpus siliculosus (multicellular brown alga),
259
Aureococcus anophagefferens (unicellular brown alga), Thalassiosira pseudonana
260
(diatom), Phaeodactylum tricornutum (diatom), and Phytophthora infestans (oomycete).
261
The situation in alveolates is more complex. Single GSK3 homologues exist in
262
Plasmodium sp. and Toxoplasma sp. with larger gene families in Paramecium tetraurelia
263
and Tetrahymena thermophila. Within excavates a single GSK3 homologue is present in
264
Leishmania sp. with two possible homologues in Trypanosoma sp. Together, these data
265
suggest a single common ancestor for GSK3. Whether the GSK3s identified in algae and
266
other protists function in osmotic stress tolerance is unknown.
267
268
Conclusions and future directions
269
GSK3 proteins have extremely diverse functions both between kingdoms and within a
270
single species. This suggests that GSK3s have been co-opted into many signalling
271
pathways throughout evolution and highlights their pivotal importance in eukaryotic
272
signalling. Plant GSK3 proteins are required for growth, development and environmental
273
stress tolerance. These are all agronomically important plant traits, making plant GSK3s
274
potential targets for future crop manipulation. So far, most functional studies have been
275
carried out in Arabidopsis or in other dicots. The exact degree of GSK3 functional
276
conservation across land plants is unknown, although rice and cotton BIN2 homologues
277
appear to have conserved function when overexpressed in Arabidopsis [39-40].
12
278
Producing dwarf plants was key to the success of the Green Revolution, and cereal semi-
279
dwarf mutants with altered BR signalling have been discovered [63-65]. Perhaps cereal
280
BIN2 homologues can be manipulated to produce semi-dwarf plants. Indeed, the
281
discovery of a GSK3 mutant [48] affecting the phosphorylation of only some substrates
282
could be a powerful tool for future development and use in crop manipulation.
283
GSK3s play a role in integrating multiple hormonal signals (BR, ABA and auxin).
284
In addition to their role in development, these hormones are involved in responding to a
285
variety of environmental stresses. Therefore, deciphering the molecular mechanism by
286
which GSK3s mediate hormone crosstalk represents an attractive way to modulate plant
287
stress tolerance.
288
Whether plant GSK3s share conserved molecular functions with GSK3s in other
289
systems is still an open question. Key unanswered questions include (i) whether plant
290
GSK3s provide a direct link between developmental cell signalling and the cytoskeleton,
291
and (ii) whether plant and animal GSK3s share any conserved protein targets and
292
upstream regulators. These should be exciting areas for future research in Arabidopsis
293
and other species.
294
295
Acknowledgements
296
We apologize to colleagues whose work could not be included because of space
297
limitations. We thank Jennifer Nemhauser for useful advice, Laura Moody for her helpful
298
comments, and three anonymous reviewers for their constructive feedback.
299
13
300
References
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
1 Cohen, P. et al. (1982) Separation and characterisation of glycogen synthase kinase 3,
glycogen synthase kinase 4 and glycogen synthase kinase 5 from rabbit skeletal muscle.
Eur. J. Biochem. 124, 21-35
2 Frame, S. and Cohen, P. (2001) GSK3 takes centre stage more than 20 years after its
discovery. Biochem. J. 359, 1-16
3 Jope, R.S. and Johnson, G.V. (2004) The glamour and gloom of glycogen synthase
kinase-3. Trends Biochem. Sci. 29, 95-102
4 Logan, C.Y. and Nusse, R. (2004) The Wnt signaling pathway in development and
disease. Annu. Rev. Cell. Dev. Biol 20, 781-810
5 Petersen, C.P. and Reddien, P.W. (2009) Wnt signaling and the polarity of the primary
body axis. Cell 139, 1056-1068
6 Harwood, A.J. et al. (1995) Glycogen synthase kinase 3 regulates cell fate in
Dictyostelium. Cell 80, 139-148
7 Kim, L. and Kimmel, A.R. (2006) GSK3 at the edge: regulation of developmental
specification and cell polarization. Curr. Drug Targets 7, 1411-1419
8 Doble, B.W. and Woodgett, J.R. (2003) GSK-3: tricks of the trade for a multi-tasking
kinase. J. Cell Sci. 116, 1175-1186
9 Jonak, C. and Hirt, H. (2002) Glycogen synthase kinase 3/SHAGGY-like kinases in
plants: an emerging family with novel functions. Trends Plant Sci. 7, 457-461
10 de la Fuente van Bentem, S. et al. (2008) Site-specific phosphorylation profiling of
Arabidopsis proteins by mass spectrometry and peptide chip analysis. J. Proteome Res. 7,
2458-2470
11 Bourouis, M. et al. (1990) An Early Embryonic Product of the Gene Shaggy Encodes
a Serine Threonine Protein-Kinase Related to the Cdc28/Cdc2+ Subfamily. Embo J. 9,
2877-2884
12 Li, J. et al. (1997) Conservation of function between mammalian and plant steroid
5alpha-reductases. Proc. Natl. Acad. Sci. U. S. A. 94, 3554-3559
13 Kim, T.W. and Wang, Z.Y. (2010) Brassinosteroid signal transduction from receptor
kinases to transcription factors. Annu. Rev. Plant Biol. 61, 681-704
14 Clouse, S.D. (2011) Brassinosteroid signal transduction: from receptor kinase
activation to transcriptional networks regulating plant development. Plant Cell 23, 12191230
15 Yang, C.J. et al. (2011) The Mechanisms of Brassinosteroids' Action: From Signal
Transduction to Plant Development. Mol. Plant 4, 588-600
16 Gudesblat, G.E. and Russinova, E. (2011) Plants grow on brassinosteroids. Current
Opinion in Plant Biology DOI: 10.1016/j.pbi.2011.05.004
17 Li, J. et al. (2001) BIN2, a new brassinosteroid-insensitive locus in Arabidopsis. Plant
Physiol. 127, 14-22
18 Li, J. and Nam, K.H. (2002) Regulation of brassinosteroid signaling by a
GSK3/SHAGGY-like kinase. Science 295, 1299-1301
19 Perez-Perez, J.M. et al. (2002) The UCU1 Arabidopsis gene encodes a
SHAGGY/GSK3-like kinase required for cell expansion along the proximodistal axis.
Dev. Biol. 242, 161-173
14
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
20 Choe, S. et al. (2002) Arabidopsis brassinosteroid-insensitive dwarf12 mutants are
semidominant and defective in a glycogen synthase kinase 3beta-like kinase. Plant
Physiol. 130, 1506-1515
21 Zhao, J. et al. (2002) Two putative BIN2 substrates are nuclear components of
brassinosteroid signaling. Plant Physiol. 130, 1221-1229
22 He, J.X. et al. (2002) The GSK3-like kinase BIN2 phosphorylates and destabilizes
BZR1, a positive regulator of the brassinosteroid signaling pathway in Arabidopsis. Proc.
Natl. Acad. Sci. U. S. A. 99, 10185-10190
23 Yin, Y. et al. (2002) BES1 accumulates in the nucleus in response to brassinosteroids
to regulate gene expression and promote stem elongation. Cell 109, 181-191
24 Ingham, P.W. et al. (2011) Mechanisms and functions of Hedgehog signalling across
the metazoa. Nat. Rev. Genet. 12, 393-406
25 Ryu, H. et al. (2007) Nucleocytoplasmic shuttling of BZR1 mediated by
phosphorylation is essential in Arabidopsis brassinosteroid signaling. Plant Cell 19,
2749-2762
26 Ryu, H. et al. (2010) Phosphorylation dependent nucleocytoplasmic shuttling of BES1
is a key regulatory event in brassinosteroid signaling. Mol. Cells 29, 283-290
27 Ginger, R.S. et al. (2000) Glycogen synthase kinase-3 enhances nuclear export of a
Dictyostelium STAT protein. Embo J. 19, 5483-5491
28 Crabtree, G.R. and Olson, E.N. (2002) NFAT signaling: choreographing the social
lives of cells. Cell 109 Suppl, S67-79
29 Peng, P. et al. (2010) A direct docking mechanism for a plant GSK3-like kinase to
phosphorylate its substrates. J. Biol. Chem. 285, 24646-24653
30 Peng, P. et al. (2008) Regulation of the Arabidopsis GSK3-like kinase
BRASSINOSTEROID-INSENSITIVE 2 through proteasome-mediated protein
degradation. Mol. Plant 1, 338-346
31 Kim, T.W. et al. (2009) Brassinosteroid signal transduction from cell-surface receptor
kinases to nuclear transcription factors. Nat. Cell Biol. 11, 1254-1260
32 Jaillais, Y. et al. (2011) Tyrosine phosphorylation controls brassinosteroid receptor
activation by triggering membrane release of its kinase inhibitor. Genes Dev. 25, 232-237
33 Lochhead, P.A. et al. (2006) A chaperone-dependent GSK3beta transitional
intermediate mediates activation-loop autophosphorylation. Mol. Cell 24, 627-633
34 Vert, G. and Chory, J. (2006) Downstream nuclear events in brassinosteroid
signalling. Nature 441, 96-100
35 Yan, Z. et al. (2009) BIN2 functions redundantly with other Arabidopsis GSK3-like
kinases to regulate brassinosteroid signaling. Plant Physiol. 150, 710-721
36 Yoo, M.J. et al. (2006) Phylogenetic diversification of glycogen synthase kinase
3/SHAGGY-like kinase genes in plants. BMC Plant Biol. 6, 3
37 Rozhon, W. et al. (2010) ASKtheta, a group-III Arabidopsis GSK3, functions in the
brassinosteroid signalling pathway. Plant J. 62, 215-223
38 De Rybel, B. et al. (2009) Chemical inhibition of a subset of Arabidopsis thaliana
GSK3-like kinases activates brassinosteroid signaling. Chem. Biol. 16, 594-604
39 Sun, Y. and Allen, R.D. (2005) Functional analysis of the BIN2 genes of cotton. Mol.
Genet. Genomics 274, 51-59
15
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
40 Koh, S. et al. (2007) T-DNA tagged knockout mutation of rice OsGSK1, an
orthologue of Arabidopsis BIN2, with enhanced tolerance to various abiotic stresses.
Plant Mol. Biol. 65, 453-466
41 Zhang, S. et al. (2009) The primary signaling outputs of brassinosteroids are regulated
by abscisic acid signaling. Proc. Natl. Acad. Sci. U. S. A. 106, 4543-4548
42 Vert, G. et al. (2008) Integration of auxin and brassinosteroid pathways by Auxin
Response Factor 2. Proc. Natl. Acad. Sci. U. S. A. 105, 9829-9834
43 Nemhauser, J.L. et al. (2004) Interdependency of brassinosteroid and auxin signaling
in Arabidopsis. PLoS Biol. 2, E258
44 Vanneste, S. et al. (2005) Cell cycle progression in the pericycle is not sufficient for
SOLITARY ROOT/IAA14-mediated lateral root initiation in Arabidopsis thaliana. Plant
Cell 17, 3035-3050
45 Dornelas, M.C. et al. (2000) Arabidopsis thaliana SHAGGY-related protein kinases
(AtSK11 and 12) function in perianth and gynoecium development. Plant J. 21, 419-429
46 Charrier, B. et al. (2002) Expression profiling of the whole Arabidopsis shaggy-like
kinase multigene family by real-time reverse transcriptase-polymerase chain reaction.
Plant Physiol. 130, 577-590
47 Tavares, R. et al. (2002) AtSKtheta, a plant homologue of SGG/GSK-3 marks
developing tissues in Arabidopsis thaliana. Plant Mol. Biol. 50, 261-271
48 Claisse, G. et al. (2007) The Arabidopsis thaliana GSK3/Shaggy like kinase AtSK3-2
modulates floral cell expansion. Plant Mol. Biol. 64, 113-124
49 Piao, H.L. et al. (2001) Constitutive over-expression of AtGSK1 induces NaCl stress
responses in the absence of NaCl stress and results in enhanced NaCl tolerance in
Arabidopsis. Plant J. 27, 305-314
50 Richard, O. et al. (2005) Organization and expression of the GSK3/shaggy kinase
gene family in the moss Physcomitrella patens suggest early gene multiplication in land
plants and an ancestral response to osmotic stress. J Mol. Evol. 61, 99-113
51 Kempa, S. et al. (2007) A plastid-localized glycogen synthase kinase 3 modulates
stress tolerance and carbohydrate metabolism. Plant J. 49, 1076-1090
52 Patade, V.Y. et al. (2011) Expression analysis of sugarcane shaggy-like kinase (SuSK)
gene identified through cDNA subtractive hybridization in sugarcane (Saccharum
officinarum L.). Protoplasma 248, 613-621
53 Chen, G.P. et al. (2003) Isolation and characterization of TaGSK1 involved in wheat
salt tolerance. Plant Science 165, 1369–1375
54 Wrzaczek, M. et al. (2007) A Proteasome-regulated glycogen synthase kinase-3
modulates disease response in plants. J. Biol. Chem. 282, 5249-5255
55 Jonak, C. et al. (2000) Wound-induced expression and activation of WIG, a novel
glycogen synthase kinase 3. Plant Cell 12, 1467-1475
56 Xiao, Y. and Mitchell, A.P. (2000) Shared roles of yeast glycogen synthase kinase 3
family members in nitrogen-responsive phosphorylation of meiotic regulator Ume6p.
Mol. Cell. Biol. 20, 5447-5453
57 Kassir, Y. et al. (2006) The Saccharomyces cerevisiae GSK-3 beta homologs. Current
Drug Targets 7, 1455-1465
58 Pereira, J. et al. (2009) Yap4 PKA- and GSK3-dependent phosphorylation affects its
stability but not its nuclear localization. Yeast 26, 641-653
16
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
59 Andoh, T. et al. (2000) Yeast glycogen synthase kinase 3 is involved in protein
degradation in cooperation with Bul1, Bul2, and Rsp5. Mol. Cell. Biol. 20, 6712-6720
60 Piao, H.L. et al. (1999) An Arabidopsis GSK3/shaggy-like gene that complements
yeast salt stress-sensitive mutants is induced by NaCl and abscisic acid. Plant Physiol.
119, 1527-1534
61 Rensing, S.A. et al. (2008) The Physcomitrella genome reveals evolutionary insights
into the conquest of land by plants. Science 319, 64-69
62 Banks, J.A. et al. (2011) The Selaginella genome identifies genetic changes associated
with the evolution of vascular plants. Science 332, 960-963
63 Chono, M. et al. (2003) A semidwarf phenotype of barley uzu results from a
nucleotide substitution in the gene encoding a putative brassinosteroid receptor. Plant
Physiol. 133, 1209-1219
64 Morinaka, Y. et al. (2006) Morphological alteration caused by brassinosteroid
insensitivity increases the biomass and grain production of rice. Plant Physiol. 141, 924931
65 Bai, M.Y. et al. (2007) Functions of OsBZR1 and 14-3-3 proteins in brassinosteroid
signaling in rice. Proc. Natl. Acad. Sci. U. S. A. 104, 13839-13844
66 Salinas, P.C. (2007) Modulation of the microtubule cytoskeleton: a role for a
divergent canonical Wnt pathway. Trends Cell Biol. 17, 333-342
67 Ong Tone, S. et al. (2010) GSK3 regulates mitotic chromosomal alignment through
CRMP4. PLoS One 5, e14345
68 Gardiner, J. and Marc, J. (2011) Arabidopsis thaliana, a plant model organism for the
neuronal microtubule cytoskeleton? J. Exp. Bot. 62, 89-97
69 Ainsworth, C. (2007) Cilia: tails of the unexpected. Nature 448, 638-641
70 Wilson, N.F. and Lefebvre, P.A. (2004) Regulation of flagellar assembly by glycogen
synthase kinase 3 in Chlamydomonas reinhardtii. Eukaryot. Cell 3, 1307-1319
17
461
Table 1. Land plant GSK3s
Arabidopsis GSK3
clade
Arabidopsis GSK3
Gene identifier
Function/remark
Refs
I
AtSK11/ASKα
At5g26750
[45,31]
AtSK12/ASKγ
At3g05840
AtSK13/ASKε
At5g14640
AtSK21/ASKη/BIN2/UCU1
AtSK22/ASKι/BIL1/AtGSK1
AtSK23/ASKζ/BIL2
AtSK31/ASKθ
At4g18710
At1g06390
At2g30980
At4g00720
AtSK32/ASKβ
AtSK41/ASKκ/AtK-1
AtSK42/ASKδ
At3g61160
At1g09840
At1g57870
Flower development/brassinosteroid
signalling
Flower development/brassinosteroid
signalling
Osmotic stress induced /brassinosteroid
signalling
Brassinosteroid signalling
Brassinosteroid signalling/salt stress
Brassinosteroid signalling
Brassinosteroid signalling/osmotic stress
induced
Flower development
Unknown
Osmotic stress induced
GSK3 gene
MsK1
MsK4
WIG
OsGSK1
TaGSK1
BIN2
SuSK
PpSK2
PpSK4
Highest similarity to
AtSK11
AtSK41
AtSK31
AtSK21/BIN2
AtSK12
AtSK21/BIN2
AtSK42
AtSK13
AtSK13
Function/remark
Pathogen response
Carbohydrate metabolism/salt stress
Wound response
Brassinosteroid signalling/salt stress
Salt stress
Brassinosteroid signalling
Salt/osmotic stress
Osmotic stress induced
Osmotic stress induced
II
III
IV
Plant species
Medicago sativa
Oryza sativa
Triticum aestivum
Gossypium sp.
Saccharum officinarum
Physcomitrella patens
[45,31]
[46,31]
[17,18,19,20]
[35,49]
[35]
[37,46]
[48]
[46]
[54]
[51]
[55]
[40]
[53]
[39]
[52]
[50]
[50]
462
18
463
Glossary
464
Alveolates: a protist lineage including Apicomplexans (unicellular parasites such as
465
Plasmodium (the malaria parasite) and Toxoplasma), marine dinoflagellates and ciliates.
466
Angiosperms: also called flowering plants. A group of land plants with reproductive
467
organs in the form of flowers, which produce seeds enclosed in capsules (or fruits). They
468
are believed to have evolved more recently (in geological terms). Modern agriculture is
469
mostly reliant on species within this group.
470
Bryophytes: Early land plants that do not have vascular tissues. They do not form
471
flowers and thus do not produce seeds. Sexual reproduction leads to the formation of
472
single-celled spores. Bryophytes are characterized by a dominant haploid gametophyte
473
life cycle. This group contain mosses, hornworts and liverworts.
474
β-catenin: is an animal protein (called Armadillo in Drosophila) containing tandem
475
copies of a protein motif called Armadillo repeat domain, which form a conserved three-
476
dimensional structure specialized for protein–protein interaction. β-catenin homologues
477
function in various processes, including intracellular signalling (e.g. Wnt signalling) and
478
cytoskeletal regulation.
479
Cellulase: A class of enzymes that can hydrolyse cellulose and degrade cell walls. They
480
have been shown to induce plant defence responses.
481
Cilia: Either motile or primary (non-motile), structurally related to flagella and are found
482
in animals. Nearly every mammalian cell has a primary cilium. Cilia are emerging as key
483
regulators of development and signalling in animal systems, particularly as homes for
484
Wnt signalling pathway components.
19
485
Cytokinesis: The process of division of the cytoplasm of a single eukaryotic cell during
486
mitosis.
487
Elicitor: Molecules produced by pathogens, or by plants in response to pathogens, that
488
activate plant defence responses.
489
Excavates: A eukaryotic kingdom consisting largely of unicellular species, some of
490
which, such as Trypanosoma and Leishmania, are disease-causing parasites.
491
Ferns: Vascular plants differing from gymnosperms and angiosperms by their mode of
492
reproduction. They reproduce via spores and constitute the earliest plants having true
493
leaves (megaphylls). This group includes horsetails (Equisetum) and lady fern
494
(Athyrium).
495
Flagella: Motile microtubule-containing cellular extensions found on unicells and on
496
motile gametes in animals, plants and protists. They act as environmental sensors and
497
control cell motility and cell–cell interactions.
498
Gymnosperms: A group of land plants that reproduce by means of 'naked seeds' exposed
499
in open structures, such as cones or leaves (in contrast to angiosperms with enclosed
500
seeds). This group includes conifers and Ginkgo.
501
Heterokonts: the eukaryotic phylum containing both unicellular and multicellular brown
502
algae, diatoms, and oomycete pathogens.
503
Lycophytes: Seedless vascular plants with primitive leaves (microphylls) having only a
504
single vascular trace (vein). Lycophytes include the resurrection plant Selaginella
505
lepidophylla and the model plant Selaginella moellendorffii, which has a sequenced
506
genome.
507
20
508
Box 1. Conserved roles for plant GSK3s in cytoskeletal regulation?
509
In animals and yeast, GSK3 functions both in cell signalling and in cytoskeletal
510
regulation. It has yet to be determined whether plant GSK3s have cytoskeletal functions.
511
However, given their instrumental roles in cell expansion and growth, it is tempting to
512
speculate that they regulate cell architecture directly.
513
Crosstalk between developmental signalling and the cytoskeleton?
514
In Caenorhabditis elegans and Drosophila, GSK3 responds to Wnt signals and controls
515
mitotic spindle orientation during cell division [66]. In mammalian cells, GSK3
516
phosphorylates a spindle microtubule-associated protein [67]. This function of GSK3
517
might be ancient given that GSK3 has several key roles in cell cycle progression in yeast
518
[57]. Thus, plant GSK3s may also be required for cell division. However, cytokinesis
519
mechanisms in plants are different from those in animal and yeast cells, suggesting that
520
novel regulatory mechanisms may also have evolved.
521
Animal GSK3 phosphorylates a plethora of microtubule-associated proteins [66],
522
some of which are conserved in plants [68]. When Wnt signals regulate GSK3 activity,
523
this leads to changes in microtubule regulation [66]. Thus, a key avenue for future studies
524
is to determine whether plant microtubule-associated proteins are regulated by GSK3,
525
and whether GSK3 provides a link between the cytoskeleton and cell growth regulation
526
by hormone/BR signalling, similar to that found in animals.
527
Flagellar functions of GSK3
528
Flagella and cilia are evolutionarily ancient organelles that regulate key cellular processes
529
[69]. The unicellular green alga Chlamydomonas is a key model system used to study the
530
molecular basis of flagellar dynamics owing to its ease of genetic manipulation.
21
531
Chlamydomonas has two motile flagella of tightly regulated length. Short-term treatment
532
with lithium, a GSK3 inhibitor, leads to elongation and paralysis of the flagella, whereas
533
longer treatments produce aflagellate cells [70]. Similarly, lithium inhibits flagellar
534
regeneration. This suggests a key role for Chlamydomonas GSK3 in regulating flagellar
535
assembly, maintenance, length and motility. RNAi-mediated knock-down of the single
536
Chlamydomonas GSK3 homologue renders cells aflagellate [70]. The active
537
(phosphorylated) form of Chlamydomonas GSK3 is enriched in the flagella (particularly
538
newly assembling flagella) and is associated with the flagellar microtubules only when
539
phosphorylated [70]. This shows that GSK3 is a key regulator of flagellar dynamics.
540
Whether this function is conserved in animals is not yet known, but given the strong
541
conservation of other flagellar protein functions between algae and animals it is an area
542
worth investigating in the future.
543
Plant cells do not have cilia, and most seed plant cells do not have flagella.
544
However, the motile gametes of the seedless land plants (bryophytes, lycophytes and
545
ferns) and some early seed plants are flagellated. These would be interesting organisms in
546
which to investigate GSK3 function.
547
548
549
Figure 1. Alignment of human GSK3 and a representative Arabidopsis GSK3,
550
AtSK21/BIN2. The kinase domain is highlighted in yellow. The conserved tyrosine
551
whose phosphorylation is required for kinase activity (Y216 in GSK3, Y200 in BIN2) is
552
highlighted in red. The conserved phosphate-binding residues that interact with prime-
553
phosphorylated substrate are highlighted in cyan. The conserved TREE domain is boxed
554
in orange. Highlighted in dark blue within this domain are the three residues that lead to
22
555
BIN2 gain-of-function alleles when point-mutated [bin2-1 and dwf12-2D (E263K); bin2-
556
2 (T261I); ucu1-1, ucu1-2 and dwf12-1D (E264K)] [18-20]. The weak allele ucu1-3
557
(P284S) is highlighted in purple. Other loss-of-function mutations (identified in [35]) are
558
highlighted in magenta (G49D, C183Y, R312K; all are in residues conserved in GSK3).
559
The kinase-dead mutation bin2 (K69R) that suppresses gain-of-function mutations is
560
highlighted in light green. Homologous residues to those identified by mutation in
561
AtSK32 [48] are marked with white arrowheads. The serine 9 residue specific to GSK3
562
and essential to the regulation of its inhibition is marked with a black arrowhead.
563
564
Figure 2. The central role of BIN2 in BR signalling and its interface with other hormone
565
pathways. In the absence of a BR signal, the active (phosphorylated) form of BIN2
566
phosphorylates BES1/BZR1 transcription factors, consequently targeting them for
567
proteasomal degradation. The BR hormone activates the BRI1 receptor and its co-
568
receptor, BAK1. This allows transphosphorylation of the two transmembrane receptors.
569
Active BRI1 phosphorylates BSK kinase, which interacts with and activates BSU1
570
phosphatases. Active BSU1 dephosphorylates BIN2, which promotes BIN2 degradation,
571
and allows unphosphorylated BES1/BZR1 to translocate to the nucleus and modulate BR
572
gene expression. BIN2 also interacts and phosphorylates the transcriptional repressor
573
ARF2. This relieves ARF2 repression on auxin-responsive promoters and leads to a
574
synergistic BR–auxin response. ABA inhibits BR signalling and enhances the
575
phosphorylation state of BES1. This action might be via a direct effect on BIN2 or an
576
upstream component.
577
23