You are on page 1of 13

Interactions between Auxin and Strigolactone in Shoot

Branching Control1[C][OA]
Alice Hayward, Petra Stirnberg, Christine Beveridge2*, and Ottoline Leyser2
University of Queensland, School of Biological Sciences, Australian Research Council Centre of Excellence for
Integrative Legume Research, Queensland 4072, Australia (A.H., C.B.); and Department of Biology, University
of York, York YO10 5YW, United Kingdom (A.H., P.S., O.L.)

In Arabidopsis (Arabidopsis thaliana), the carotenoid cleavage dioxygenases MORE AXILLARY GROWTH3 (MAX3) and MAX4
act together with MAX1 to produce a strigolactone signaling molecule required for the inhibition of axillary bud outgrowth.
We show that both MAX3 and MAX4 transcripts are positively auxin regulated in a manner similar to the orthologous genes
from pea (Pisum sativum) and rice (Oryza sativa), supporting evolutionary conservation of this regulation in plants. This
regulation is important for branching control because large auxin-related reductions in these transcripts are associated with
increased axillary branching. Both transcripts are up-regulated in max mutants, and consistent with max mutants having
increased auxin in the polar auxin transport stream, this feedback regulation involves auxin signaling. We suggest that both
auxin and strigolactone have the capacity to modulate each other’s levels and distribution in a dynamic feedback loop required
for the coordinated control of axillary branching.

Shoot branching is dependent on both the formation axillary bud outgrowth (Turnbull et al., 2002; Sorefan
of axillary meristems in the axils of leaves and the et al., 2003; Booker et al., 2004, 2005; Auldridge et al.,
precise control of their outgrowth by genetically, hor- 2006).
monally, and environmentally regulated signals (for This mobile signal appears to be a strigolactone or a
review, see Schmitz and Theres, 2005; Dun et al., 2006; derivative (Gomez-Roldan et al., 2008; Umehara et al.,
Ongaro and Leyser, 2007). Mutations in the Arabidop- 2008). Strigolactones are carotenoid-derived terpenoid
sis (Arabidopsis thaliana) MORE AXILLARY GROWTH lactones previously shown to stimulate mycorrhizal
(MAX) genes and orthologous genes in garden pea hyphal branching and seed germination of the plant
(Pisum sativum), rice (Oryza sativa), and petunia (Petu- parasitic weeds Striga and Orobanche (Cook et al., 1972;
nia hybrida) all lead to increased branching (Beveridge, Akiyama et al., 2005; Humphrey and Beale, 2006). ccd7
2000; Stirnberg et al., 2002; Sorefan et al., 2003; Ishikawa and ccd8 mutants of rice and pea are deficient in
et al., 2005; Snowden et al., 2005; Zou et al., 2005, 2006; strigolactones, and the strigolactone analog, GR24
Johnson et al., 2006; Arite et al., 2007; Simons et al., (Johnson et al., 1981), can restore branching inhibition
2007). Two of these genes, MAX3 (AtCCD7) and MAX4 in these mutants as well as in max1, max3, and max4
(AtCCD8), encode plastid-targeted carotenoid cleav- (Gomez-Roldan et al., 2008; Umehara et al., 2008). In
age dioxygenases (CCDs) that together with the cy- the shoot, strigolactone signal transduction probably
tochrome P450 family member, MAX1, control the involves the F-box Leu-rich repeat protein, MAX2
production of an upwardly mobile signal that inhibits (Stirnberg et al., 2002, 2007; Booker et al., 2005). Ac-
cordingly, branching in max2 plants is not repressed by
1
This work was supported by the Australian Research Council
GR24. F-box proteins function in Skp1-Cul1/Cdc53-
Centre of Excellence for Integrative Legume Research, by the Bio- F-box (SCF) E3 ubiquitin ligase complexes, which
technology and Biological Sciences Research Council, and by an target proteins for ubiquitination, often triggering
Australian Postgraduate Award, a Travelling Fellowship from the subsequent degradation by the 26S proteosome (for
Company of Biologists (Development), and a University of Queens- review, see Petroski and Deshaies, 2005).
land Graduate School Research Travel Award to A.H. A second hormone central to branching control is
2
These authors contributed equally to the article. auxin, which was first shown to inhibit branching over
* Corresponding author; e-mail c.beveridge@uq.edu.au. seven decades ago (Thimann and Skoog, 1934). Auxin
The author responsible for distribution of materials integral to the also signals via SCF-mediated targeted protein degra-
findings presented in this article in accordance with the policy dation. The F-box protein, TRANSPORT INHIBITOR
described in the Instructions for Authors (www.plantphysiol.org) is:
Christine Beveridge (c.beveridge@uq.edu.au).
RESPONSE1 (TIR1), and closely related family mem-
[C]
Some figures in this article are displayed in color online but in bers (AFBs) act as auxin receptors (Dharmasiri et al.,
black and white in the print edition. 2005a, 2005b; Kepinski and Leyser, 2005). Auxin bind-
[OA]
Open Access articles can be viewed online without a sub- ing stabilizes the interaction between TIR1/AFBs and
scription. members of the Aux/IAA family of transcriptional
www.plantphysiol.org/cgi/doi/10.1104/pp.109.137646 repressors (Tan et al., 2007). This induces Aux/IAA
400 Plant PhysiologyÒ, September 2009, Vol. 151, pp. 400–412, www.plantphysiol.org Ó 2009 American Society of Plant Biologists
Downloaded from www.plantphysiol.org on June 24, 2014 - Published by www.plant.org
Copyright © 2009 American Society of Plant Biologists. All rights reserved.
Auxin and Strigolactone Interactions in Branching

degradation, allowing auxin-mediated up-regulation expression in a manner typical of many hormone


of transcription (Gray et al., 2001; Tiwari et al., 2001; biosynthetic pathways (Wang et al., 2002; Yamaguchi,
Zenser et al., 2001; Dharmasiri and Estelle, 2002). As a 2008). However, systemic feedback regulation requires
homeostatic mechanism, many Aux/IAA genes are a signal in addition to strigolactone, since strigolac-
rapidly transcriptionally induced by auxin in an tones move upward while systemic feedback signaling
SCFTIR1/AFB-dependent manner (Park et al., 2002). An- has been shown to move down the plant (Foo et al.,
other protein, AUXIN RESISTANT1 (AXR1), is neces- 2005; Johnson et al., 2006). In pea and Arabidopsis,
sary for proper SCF function by facilitating the feedback signaling from the shoot also down-regulates
conjugation of the ubiquitin-like protein, RUB1, to the levels of the branching promoter, cytokinin, in the
the Cullin subunit (Leyser et al., 1993; Wu et al., 2000; root xylem sap (Foo et al., 2007).
Schwechheimer et al., 2001; del Pozo et al., 2002). In Arabidopsis, mutations in the MAX genes result
Mutations in AXR1 cause changes in SCFTIR1/AFB- in increased stem conductivity for auxin and increased
dependent auxin-responsive gene expression and sub- expression of several auxin transporters, including
sequent defects in downstream auxin responses. AXR1 some of the PIN family of auxin exporters (Bennett
also appears to regulate SCF signaling in photomor- et al., 2006; Lazar and Goodman, 2006; Brewer et al.,
phogenesis and jasmonate responses (Schwechheimer 2009). The stems of max mutants have increased ex-
et al., 2002; Tiryaki and Staswick, 2002). Phenotypes pression from the synthetic auxin-responsive pro-
conferred by axr1 include increased branching, agravi- moter DR5, suggestive of increased auxin content.
tropic root growth, fewer lateral roots and root hairs, Similarly, increased auxin levels have been observed in
and reduced fertility due to poor stamen elongation strigolactone mutants of rice (Arite et al., 2007) and
(Lincoln et al., 1990). sometimes, but not always, in pea (for review, see Dun
Buds of strigolactone pathway mutants are resistant et al., 2006). Therefore, high auxin content in the stems
to inhibition by apically applied auxin, suggesting that of max mutants is a good candidate for mediating
strigolactones are required for this inhibition (Bever- feedback regulation of the MAX pathway. Low strigo-
idge, 2000; Sorefan et al., 2003; Bennett et al., 2006). lactone would lead to increased auxin content, which
Additionally, GR24 can inhibit branching in auxin- in turn would up-regulate strigolactone synthesis. In
depleted decapitated pea plants and in auxin signaling pea, both auxin and another long-distance feedback
mutants, including axr1 and the tir1afb1afb2afb3 qua- signal are suggested to be involved in feedback (for
druple mutant (Brewer et al., 2009). In pea, the genes review, see Dun et al., 2006). Therefore, auxin and
orthologous to MAX3 and MAX4, RAMOSUS5 strigolactone signaling could interact in interlocking
(RMS5) and RMS1, are highly auxin regulated, with feedback loops, with each hormone having the poten-
transcripts greatly diminished or slightly increased by tial to regulate the levels of the other. Here, we
treatments shown to deplete or enhance auxin, respec- describe the results of experiments aimed at further
tively (Foo et al., 2005; Johnson et al., 2006). Auxin investigating the auxin-strigolactone interaction in
regulation was also shown for the rice orthologs, Arabidopsis and exploring its functional significance.
HIGH TILLERING DWARF1 and DWARF10 (D10;
Zou et al., 2006; Arite et al., 2007). These data suggest
that the auxin regulation of strigolactone synthesis
genes could represent a conserved and important RESULTS
mode of auxin action in branching control. In Arabi- MAX3 and MAX4 Expression in the Shoot Is
dopsis, the response of MAX3 and MAX4 transcripts Up-Regulated by Auxin in an AXR1-Dependent Manner
to auxin depletion has not been assessed in detail. A
MAX4 promoter:GUS fusion was responsive to auxin Auxin regulation of RMS1/MAX4 and RMS5/MAX3
treatments, particularly in the root elongation zone in stems has been detected in pea but not in Arabi-
and in hypocotyls, and this was dependent on AXR1 dopsis (Bainbridge et al., 2005; Foo et al., 2005; Johnson
(Bainbridge et al., 2005). In contrast, up-regulation of et al., 2006). To determine if auxin up-regulation of
GUS by auxin was not detected in shoot tissues. MAX3 and MAX4 can be detected in the shoot using
In strigolactone pathway mutants of pea, RMS1 and the more sensitive and quantitative quantitative real-
RMS5 transcription is feedback up-regulated (Foo time PCR (qRT-PCR), and whether this effect is AXR1
et al., 2005; Johnson et al., 2006). This was also shown dependent, a dose response of MAX3 and MAX4
for the MAX4 ortholog in rice and petunia (Snowden expression to the natural auxin indole-3-acetic acid
et al., 2005; Arite et al., 2007; Simons et al., 2007). A (IAA) was undertaken in the basal cauline internode of
comparatively minor increase in MAX4 transcription intact axr1-3 and wild-type plants. The INDOLE-
has been reported in Arabidopsis max2 hypocotyls 3-ACETIC ACID INDUCIBLE1 (IAA1) gene known to
(Bainbridge et al., 2005). In pea, grafting experiments be up-regulated by auxin via the AXR1/TIR1 pathway
indicate that feedback up-regulation of RMS1 and (Park et al., 2002; Yang et al., 2004) was analyzed as a
RMS5 occurs both locally and systemically (Foo et al., control.
2005; Johnson et al., 2006; for review, see Dun et al., In pea, 3 and 19 mM IAA applied in lanolin to
2006). Local feedback regulation could be mediated by decapitated plants can substantially enhance total
a direct effect of strigolactone signaling on RMS gene stem auxin content and slightly but significantly up-
Plant Physiol. Vol. 151, 2009 401
Downloaded from www.plantphysiol.org on June 24, 2014 - Published by www.plant.org
Copyright © 2009 American Society of Plant Biologists. All rights reserved.
Hayward et al.

regulate RMS1 and RMS5 expression (Foo et al., 2005; Young expanding leaves at the shoot apex are im-
Johnson et al., 2006). When tested here, 1 and 19 mM portant sources of auxin (Thimann and Skoog, 1934;
IAA, but not 100 mM or below, significantly increased Ljung et al., 2001). In 5-week-old wild-type plants, re-
MAX3 and MAX4 transcript levels in the basal cauline moving these sources by decapitation, or applying the
internode of intact plants by at least 10-fold at 3 h (t test auxin polar transport inhibitor 1-N-naphthylphthalamic
P , 0.01; Fig. 1). These genes behaved similarly to acid (NPA) in lanolin, significantly depleted MAX3
IAA1, although IAA1 was slightly more responsive to transcripts in tissues below (Fig. 2A; all P , 0.05).
IAA addition. For all genes, the response to IAA was Specifically, these treatments reduced MAX3 transcripts
clearly diminished in the axr1-3 mutant (P , 0.01– in the primary bolting stem by 200- to 400-fold after
0.05). In untreated plants, expression levels were 24 h. In hypocotyls and roots, the greatest reductions
slightly reduced in axr1-3 relative to the wild type in MAX3 transcript occurred when plants were de-
for MAX4 (P , 0.05) and IAA1 (P , 0.01), while this capitated, including removal of young rosette leaves,
was not significant for MAX3. Together, these data or when NPA was applied below these leaves around
suggest that, like IAA1, MAX3 and MAX4 transcripts the top of the hypocotyls (all P , 0.05). Qualitatively
are induced by auxin in internode tissue in an AXR1- similar, but generally quantitatively smaller, changes
dependent manner. As is typical with auxin responses were seen for MAX4 transcripts (data not shown).
in axr1 mutants, the auxin dose-response curve is The reductions in MAX3 and MAX4 transcript levels
shifted such that high auxin doses successfully induce in hypocotyls following NPA treatment and decapita-
a response similar to that observed with lower doses in tion were generally not as great in max mutants as in
the wild type. wild-type plants (Fig. 2, B and C). This was particu-
larly obvious for max2, the most severe max mutant,
MAX3 and MAX4 Expression Is Reduced by Auxin where there was no significant difference in transcript
Depletion Treatments levels between intact and decapitated plants (P $ 0.1).
A similar effect was seen for IAA1, which while
In pea, the largest changes in RMS5 and RMS1 relatively less responsive to decapitation in general
expression occur following treatments that cause shows a significant reduction in transcript levels (P ,
auxin depletion (Foo et al., 2005; Johnson et al., 0.05) except in max1 and max2 (Fig. 2D), possibly
2006). Thus, to investigate this for MAX3 and MAX4, reflecting the increased auxin content/signaling prop-
transcript levels were quantified following treatments erties of these mutants (Bennett et al., 2006). In decap-
predicted to deplete auxin in a number of tissues. itated plants of all genotypes, transcript loss was
prevented by the application of 19 mM IAA to the
decapitation site.

Auxin-Related Down-Regulation of MAX3 and MAX4


Expression Can Activate Branching
Grafting studies revealed that wild-type roots can
restore the branching in max3 and max4 shoots to wild-
type levels while max3 roots cannot rescue max4 shoots
and visa versa (Turnbull et al., 2002; Sorefan et al.,
2003; Booker et al., 2005). This indicates that MAX3
and MAX4 are required in the same tissue for strigo-
lactone production and that their product can move
from the root to the shoot to inhibit branching. Previ-
ously, roots of the axr1-3 mutant were also shown to
restore the branching in max4 shoots to wild-type
levels, suggesting that the axr1-3 roots of these grafts
are not strigolactone deficient (Sorefan et al., 2003;
Bainbridge et al., 2005). These and other data (Bennett
et al., 2006) suggest a MAX-independent role for auxin
signaling in branching suppression, but they do not
rule out a role for AXR1/TIR1-mediated regulation of
MAX3 and MAX4 transcription in modulating branch-
Figure 1. Auxin dose response of MAX3, MAX4, and IAA1 expression
ing levels. To investigate this possibility, we turned to
in wild-type (WT) and axr1-3 basal cauline internodes. IAA was applied
in a lanolin ring around the primary bolt of 5-week-old plants 15 mm
the bodenlos-2 (bdl-2) mutant. This is a semidominant
above the base, and the cauline tissue below the application site was auxin response mutant that confers defective primary
collected 3 h post application. Expression is shown relative to wild-type root formation in the embryo and greatly increased
controls for each gene; data are means of two biological pools of six shoot branching in the adult plant, particularly obvi-
plants 6 SE. Asterisks indicate significant differences between axr1 and ous in heterozygous plants (Fig. 3). The bdl-2 allele was
the wild type. isolated from an ethyl methanesulfonate-mutagenized
402 Plant Physiol. Vol. 151, 2009
Downloaded from www.plantphysiol.org on June 24, 2014 - Published by www.plant.org
Copyright © 2009 American Society of Plant Biologists. All rights reserved.
Auxin and Strigolactone Interactions in Branching

Figure 2. A, MAX3 expression in re-


sponse to auxin-depleting treatments
in various tissues of 5-week-old wild-
type (WT) plants relative to intact hy-
pocotyls. Plants were intact, treated
with 3 mg mL21 NPA in a lanolin ring
around the primary bolt 15 mm above
the base (NPA-treated bolt) or around
the top of the hypocotyl (NPA-treated
hypocotyl), or decapitated (Decap) 15
mm above the base of the primary
inflorescence with or without young
rosette leaves removed. “Bolt” refers
to the basal 1 cm of primary cauline
stem below the treatment site. B to
D, MAX3 (B), MAX4 (C), and IAA1
(D) expression in response to auxin-
depleting treatments in the hypocotyls
of 3-week-old wild-type and max mu-
tant plants. Plants were intact, treated
with 3 mg mL21 NPA in a lanolin ring
around the top of the hypocotyl, or
decapitated by removing the entire
primary bolt and youngest leaves with
or without 19 mM IAA applied to the
decapitation site. Tissues were col-
lected 24 h after treatments. Data are
means of two to three biological pools
of seven to 25 plants 6 SE. Asterisks
indicate significant differences from
intact plants per genotype/tissue.

population in the ecotype Columbia (Col-0) back- MAX4 in the hypocotyls and inflorescence stems (data
ground (Stirnberg et al., 2002). It has a Pro-74-to-Ser not shown); however, IAA12, MAX3, and MAX4 all
amino acid substitution in domain II of the Aux/IAA showed highest differential expression in the hypo-
transcriptional repressor IAA12 identical to that de- cotyl xylem, while IAA1 was expressed most highly in
scribed for bdl-1, which is in the Landsberg erecta callus tissue and the lateral root cap.
genetic background (Hamann et al., 1999, 2002). This To investigate whether low MAX3 and MAX4 ex-
substitution results in auxin-resistant stabilization pression in bdl-2 mutants correlates with insufficient
of the IAA12/BDL protein and thus likely consti- graft-transmissible branching suppression, as ex-
tutive repression of target auxin-up-regulated genes pected if strigolactone levels have been affected, we
(Hamann et al., 2002; Dharmasiri et al., 2005b). The carried out grafting between bdl-2 heterozygotes,
bdl-2 mutation caused increased bud outgrowth from max4-1, and the wild type. If bdl-2 mutants are strigo-
the rosette in both homozygotes and heterozygotes. lactone deficient, two results were expected. First, bdl-2
Shoot height was far more reduced in homozygotes, mutant roots should not rescue branching in max4-1
which were severely dwarfed, than in heterozy- shoots; second, branching in bdl-2 shoots should
gotes, which with moderately shorter stature and in- be rescued by wild-type but not max4-1 rootstocks.
creased branching somewhat resembled max mutants Self-grafted bdl-2 heterozygous mutants produced sig-
(Fig. 3B). nificantly more rosette branches than self-grafted
Transcript abundance of MAX3 and MAX4 was max4-1 mutants (P , 0.01; Fig. 3E), and both mutants
reduced by approximately 3 orders of magnitude in branched more than self-grafted wild-type plants
bdl-2 homozygotes and also to a lesser extent in (P , 0.001). As expected, branching in max4-1 scions
heterozygotes (Fig. 3D). IAA1 expression was equal was completely restored to wild-type levels by graft-
to, or higher than, wild-type expression in bdl-2, ing to wild-type rootstocks (Sorefan et al., 2003; Fig. 3,
suggesting distinct Aux/IAA regulation of these E and F). However, bdl-2 heterozygote rootstocks did
genes. To examine this hypothesis further, we searched not restore the branching in max4-1 shoots, suggesting
for coexpression of IAA1, IAA12, MAX3, and MAX4 that BDL affects root strigolactone production. Fur-
using the Genevestigator V3 microarray database thermore, wild-type rootstocks inhibited branching in
(Zimmermann et al., 2004; Hruz et al., 2008). IAA12 heterozygote bdl-2 scions by over 50% relative to bdl-2
and IAA1 were both corepresented with MAX3 and heterozygote self grafts (P , 0.001) and compared with
Plant Physiol. Vol. 151, 2009 403
Downloaded from www.plantphysiol.org on June 24, 2014 - Published by www.plant.org
Copyright © 2009 American Society of Plant Biologists. All rights reserved.
Hayward et al.

Figure 3. Rosette branching and gene


expression in bdl mutants. A to C,
Representative rosette phenotypes of
5-week-old wild-type plants (WT; A),
bdl-2 heterozygotes (B), and bdl-2 ho-
mozygotes (C). Blue arrows indicate
the primary inflorescence stem; white
arrows indicate example axillary ro-
sette branches/buds. D, Relative ex-
pression levels of MAX3, MAX4, and
IAA1 in the basal cauline internodes of
5-week-old bdl-2 homozygotes and
heterozygotes. Data are means of two
biological pools of eight to 10 plants 6
SE; asterisks indicate significant differ-
ences from the wild type. E, Number of
rosette branches 5 mm or longer in
reciprocally grafted wild-type plants,
max4-1 mutants, and bdl-2 heterozy-
gotes 6 SE; n = 4 to 10. F, Photographs
of representative max4-1 grafted plants
(scion/rootstock). G, Number of rosette
branches 5 mm or longer per rosette
leaf in wild-type plants, max2-1 and
max4-1 mutants, and bdl-2 heterozy-
gotes with or without GR24; n = 13 to
21. Data are means 6 SE; asterisks in-
dicate significant inhibition of branch-
ing by GR24. [See online article for
color version of this figure.]

only 20% inhibition by max4-1 rootstocks (P , 0.05). Auxin Signaling Contributes to Feedback Up-Regulation
These data support the hypothesis that increased of MAX3 and MAX4 Expression in max Mutants
branching in bdl-2 mutants is partly due to insufficient
strigolactones but also that there is a strigolactone- Sixfold to 10-fold increases in MAX3 and MAX4
independent component to the increased branching expression were observed in hypocotyls of all max
observed in bdl-2 heterozygous plants. mutants relative to the wild type, indicative of feed-
We also determined the ability of exogenous strigo- back up-regulation (Fig. 2, B and C). Given that auxin
lactone to inhibit branching in bdl-2. The synthetic levels and/or signaling are increased in max mutant
strigolactone analog, GR24, was applied to rosette stems (Bennett et al., 2006), MAX3 and MAX4 tran-
axils every 2 to 3 d as described by Gomez-Roldan scription is auxin regulated in an AXR1-dependent
et al. (2008). This reduced rosette branching in max4-1 manner (Bainbridge et al., 2005; this study), and auxin
and bdl-2 heterozygotes by a similar degree (39% and is downwardly mobile, it was hypothesized that this
34%, respectively; P , 0.001; Fig. 3G), but branching feedback up-regulation may be in part mediated by
remained increased relative to the wild type for both auxin. To investigate this possibility, we determined
genotypes, particularly in bdl-2. Branching was un- the role of AXR1 in feedback signaling using the axr1-
changed in both the wild type and, as expected for a 3max2-1 double mutant (Bennett et al., 2006).
putative strigolactone signal transduction mutant, in In the basal cauline internodes of 5-week-old plants,
max2-1 (Booker et al., 2005; Stirnberg et al., 2007; MAX3, MAX4, and IAA1 transcripts accumulated to
Gomez-Roldan et al., 2008; Umehara et al., 2008). significantly higher levels in max2-1 mutants than in
404 Plant Physiol. Vol. 151, 2009
Downloaded from www.plantphysiol.org on June 24, 2014 - Published by www.plant.org
Copyright © 2009 American Society of Plant Biologists. All rights reserved.
Auxin and Strigolactone Interactions in Branching

wild-type plants (P , 0.01; Fig. 4A). In the axr1-3max2-1 sion in max2-1 hypocotyls is not related to auxin;
background, the expression of all genes in this tissue however, like MAX3 and MAX4, IAA1 and additional
was reduced to levels not statistically significantly auxin-responsive Aux/IAA genes (IAA5, IAA19, and
different from the wild type or the single axr1-3 mu- IAA29; data not shown) showed increased expression
tant. Thus, axr1 is required for feedback up-regulation in max2-1 that was not significantly affected by the
in the shoot. In the hypocotyls of the same plants, the axr1-3 mutation.
up-regulation of MAX3 and MAX4 in max2-1 was less In the hypocotyls of vegetative 2-week-old max2-1
than in cauline tissue (significant at P , 0.05; Fig. 4B), plants, MAX3, MAX4, and IAA1 transcripts were again
and this was not abolished in the axr1-3 background. slightly but significantly up-regulated (P , 0.05; Fig.
This might indicate that the observed elevated expres- 4C). At this developmental stage, the axr1-3 back-
ground prevented the up-regulation of IAA1, but not
MAX3 or MAX4, in max2-1 hypocotyls. This could
suggest auxin-independent feedback specific to MAX3
and MAX4. However, we cannot rule out an auxin-
dependent relationship, because NPA treatment of
these plants abolished the up-regulation of MAX3
and MAX4 in axr1-3max2-1 relative to axr1-3 (Fig. 4D),
and our results with bdl-2 above suggest that differ-
ences in the auxin responsiveness of these genes could
be due to different Aux/IAAs involved in their regu-
lation.

Graft Transmissibility of Feedback Signals


To test if feedback regulation of MAX3 and MAX4
expression can occur over long distances, as shown in
pea (Foo et al., 2005; Johnson et al., 2006), we recipro-
cally grafted max2-1 and wild-type plants.

Feedback in the Shoot


As shown for pea, a local feedback effect was noted
in the shoot, as evidenced by the increased expression
of MAX3, MAX4, and IAA1 in the shoots of max2/
wild-type (scion/rootstock) grafts compared with
wild-type/wild-type grafts (Fig. 5A). For all grafts,
the expression of all genes in the shoot was dependent
only on the shoot genotype; therefore, similar to re-
sults in pea, no upwardly mobile feedback signaling
was detected (Foo et al., 2005; Johnson et al., 2006).

Feedback in the Root


The expression of MAX3, MAX4, and IAA1 was
higher in the rootstocks of wild-type/max2 grafts
relative to wild-type/wild-type grafts, supporting a
local feedback effect also in the root (Fig. 5B). How-
ever, MAX3, MAX4, and IAA1 expression was not
increased in the rootstocks of max2/wild-type grafts
relative to wild-type/wild-type grafts. Therefore, in
contrast with results in pea (Foo et al., 2005), a down-
wardly mobile feedback signal from max2-1 shoots,
able to up-regulate gene expression in wild-type roots,
Figure 4. Expression of MAX3, MAX4, and IAA1 in wild-type (WT),
was not detected. Nonetheless, the increased expres-
axr1-3, max2-1, and axr1-3max2-1 plants. A and B, Expression in the
basal cauline internodes (A) and hypocotyls (B) of 5-week-old plants. C,
sion of MAX3 and MAX4, but not IAA1, in max2
Expression in the hypocotyls of vegetative 2-week-old plants; samples rootstocks was reduced by approximately 40% when
marked with “N” were treated with 3 mg mL21 NPA around the top of grafted to a wild-type shoot compared with a max2
the hypocotyls (note log scale). Data are means of two biological pools shoot (Fig. 5B; P , 0.05). This is suggestive of some
of five to 10 plants 6 SE. Asterisks indicate a significant effect of axr1 on long-distance effect on MAX3 and MAX4 expression
feedback in max2. in the rootstock depending on the shoot genotype.
Plant Physiol. Vol. 151, 2009 405
Downloaded from www.plantphysiol.org on June 24, 2014 - Published by www.plant.org
Copyright © 2009 American Society of Plant Biologists. All rights reserved.
Hayward et al.

cots and dicots by regulating the transcription of


MAX3 (CCD7) and MAX4 (CCD8) orthologs (Sorefan
et al., 2003; Foo et al., 2005; Johnson et al., 2006; Zou
et al., 2006; Arite et al., 2007). Here, we reveal that
auxin positively regulates both MAX3 and MAX4
expression in Arabidopsis more extensively than pre-
viously thought. This further supports conservation,
and hence possible functional significance, of this
regulation throughout evolution.
In the shoot, auxin enhanced MAX3 and MAX4
transcript levels within 3 h, and this was AXR1
dependent (Fig. 1). However, high concentrations
(.1 mM) of IAA in lanolin were required to elicit this
response, even for the IAA1 gene, which in etiolated
seedlings can be induced by as little as 1 mM IAA in
solution over a similar time frame (Abel et al., 1995).
This may suggest reduced permeability or sensitiv-
ity of intact light-grown cauline tissue to IAA ap-
plied in lanolin. Similarly to pea, rapid and large
reductions in MAX3 and MAX4 transcript levels
were observed following treatments predicted to
deplete endogenous auxin, and these reductions
were prevented by auxin addition (Fig. 2). Together,
these data imply that the basal levels of MAX3 and
MAX4 transcripts in intact wild-type plants are
substantially maintained by auxin. Consistent with
this idea, the levels of MAX3 and MAX4 transcript in
the auxin response mutant axr1-3 were in general
Figure 5. Expression of MAX3, MAX4, and IAA1 in the shoot (A) and lower than in the wild type, although these differ-
the rootstock (B) of grafted wild-type (WT) and max2-1 plants. Expres- ences were not always statistically significant (Figs.
sion was measured in vegetative plants 2 weeks after transfer to soil; 1 and 4). This is likely due in part to reduced
values are shown relative to wild-type/wild-type (scion/rootstock) feedback inhibition on auxin synthesis in the axr1-3
rootstocks. Data are means of two biological pools of five to eight mutant, resulting in increased auxin levels (Romano
plants 6 SE. Bars with asterisks are significantly different from each
et al., 1995), which may partially compensate for
other due to a long-distance effect.
reduced auxin sensitivity. Furthermore, AXR1 acts
partly redundantly with an additional AXR1-like
As growing shoot apices export auxin, it was con- protein, AXL1 (Dharmasiri et al., 2007), and the axr1-3
sidered that auxin, or additional, downwardly mobile allele used in this work is not a null allele. AXR1-
feedback signals may be enhanced by the presence of dependent auxin regulation of MAX3 and MAX4 is
actively growing branches in max mutants. Thus, consistent with the existence of auxin-response cis-
grafts were repeated and grown until plants had elements in the promoter regions of these genes,
primary bolting stems 15 to 30 cm in length and including four and two copies, respectively, of the
rosette branches. MAX3, MAX4, and IAA1 were TGTCTC ARF-binding motif (Ulmasov et al., 1999)
up-regulated in the rootstocks of branching max2 within 3 kb upstream.
self-grafts by a similar degree as in vegetative max2 In contrast to the mild effect of the axr1-3 mutant on
self-grafts and again were not up-regulated in wild- MAX3 and MAX4 transcript levels, the highly
type rootstocks grafted to max2 shoots (data not branched bdl-2 mutant (Fig. 3), expressing the
shown). Therefore, while there appears to be both a IAA12 transcriptional repressor that is resistant to
local and a long-distance feedback effect of strigolac- AXR1/TIR1-mediated destabilization (Hamann et al.,
tone signaling on MAX3 and MAX4 transcript abun- 2002; Dharmasiri et al., 2005b), had transcript levels
dance in Arabidopsis, it appears that, in contrast to similar to those of plants treated to deplete endoge-
pea, this long-distance effect is relatively minor. nous auxin (Figs. 2 and 3D). Thus, an attractive
hypothesis, also suggested by Brewer et al. (2009), is
that substantial auxin depletion results in increased
DISCUSSION branching at least in part by reducing strigolactone
Auxin Can Affect Strigolactone-Mediated Branch synthesis. Consistent with this, branching in bdl-2
Inhibition via MAX3 and MAX4 Regulation shoots appears to involve strigolactone deficiency.
First, shoot branching in max4 mutants can be re-
Previous analyses have shown that auxin may in- stored to wild-type levels by grafting to wild-type
teract with the strigolactone pathway in both mono- rootstocks but not bdl-2 rootstocks (Sorefan et al.,
406 Plant Physiol. Vol. 151, 2009
Downloaded from www.plantphysiol.org on June 24, 2014 - Published by www.plant.org
Copyright © 2009 American Society of Plant Biologists. All rights reserved.
Auxin and Strigolactone Interactions in Branching

2003; Bainbridge et al., 2005; Fig. 3E), suggesting that with max4 roots could involve differences in cytoki-
bdl-2 rootstocks do not produce sufficient strigolac- nin levels, caused by auxin insensitivity of bdl-2
tone. Second, wild-type rootstocks rescued over half root and shoot tissues. However, while local produc-
of the branching in bdl-2 shoots, while strigolactone- tion of cytokinin in shoot tissues can affect branch-
deficient max4-1 rootstocks did not, suggesting this ing, the importance of root-derived cytokinin to
strigolactone deficiency may be at least partly re- axillary branching remains controversial (Chaudhury
sponsible for the increased shoot branching in bdl et al., 1993; Beveridge et al., 1997a; Faiss et al., 1997;
(Fig. 3, E–G). Consistent with this idea, GR24 sup- McKenzie et al., 1998; Catterou et al., 2002; Schmülling,
pressed branching by a similar degree in bdl-2 and 2002; Sakamoto et al., 2006).
max4-1. An interaction between BDL/IAA12 and
MAX3 and MAX4 is consistent with their preferential The Mechanisms of Strigolactone Feedback Regulation
expression in the xylem and the localization of a BDL: in Arabidopsis
GUS fusion protein in vascular tissues (Zimmermann
et al., 2004; Dharmasiri et al., 2005b; Hruz et al., 2008). We show that both MAX3 and MAX4 transcripts are
Although strigolactone appears to be involved in the up-regulated in all max mutants, suggesting feedback
shoot-branching phenotype of bdl-2 mutants, it is control (Fig. 2, B and C). The degree of up-regulation
unlikely to play a role in the embryonic phenotype varies between tissues and is of a similar magnitude to
(Hamann et al., 1999, 2002; Hardtke et al., 2004), since that for orthologous genes in rice and petunia, being
mutants for the strigolactone pathway do not show substantially less than that for RMS1 in pea, where
similar pleiotropic abnormalities. transcripts can accumulate to more than 1,000-fold in
Taken together, our results suggest that auxin- rms4 mutants (Foo et al., 2005; Snowden et al., 2005;
regulated MAX expression can be blocked by stabi- Arite et al., 2007; Simons et al., 2007; Umehara et al.,
lization of BDL/IAA12, and this can modulate the 2008). There are at least three not mutually exclusive
degree of branching. Conversely, consistent with max4 potential mechanisms for the up-regulation of MAX3
shoot branching being rescued by grafting to axr1 roots and MAX4 gene expression in max mutants: (1) it is a
(Bainbridge et al., 2005), the mild down-regulation direct effect of the strigolactone signaling pathway; (2)
of MAX3 and MAX4 transcripts observed in axr1-3 it is associated with the demonstrated effect of auxin
mutant roots may not be sufficient for strigolactone- on MAX3 and MAX4 expression, coupled with the
dependent modulation of branching. The increased effect of the MAX pathway on auxin transport/levels
branching observed in axr1-3 mutants, the additive and thus auxin distribution; and (3) it is associated
bud auxin-insensitivity and branching phenotype of with an additional downwardly mobile strigolactone-
axr1-3max double mutants relative to the single mu- regulated signal, as proposed for pea (Foo et al., 2005;
tants, and the normal levels of auxin transport in Dun et al., 2006).
axr1-3 mutants also point to mutually independent
roles for MAX and auxin signaling in suppressing Direct Effects of Strigolactone Signaling on Strigolactone
branching (Bainbridge et al., 2005; Bennett et al., 2006). Biosynthesis Gene Expression
Although some effects of axr1 mutants may not be
related solely to auxin signaling (Schwechheimer et al., The best evidence for a direct feedback effect of
2002; Tiryaki and Staswick, 2002), an additional role strigolactone signaling comes from studies of rice
for auxin signaling in branching is supported here roots, where up-regulated D10 expression in strigo-
by the fact that branching in bdl-2 shoots was par- lactone pathway mutants can be restored to wild-type
tially suppressed by max4-1 roots compared with bdl levels by strigolactone application in the biosynthesis
self-grafts, suggesting that some of the branching in mutant but not the signaling mutant (Umehara et al.,
bdl-2 cannot be explained by strigolactone deficiency 2008). However, this was determined 24 h after strigo-
(Fig. 3). lactone application and thus remains of limited value
An additional mode of auxin action is likely to in distinguishing a direct effect from one mediated
involve down-regulated synthesis of the branching by another signal such as auxin, considering that
promoter cytokinin (Wickson and Thimann, 1958, auxin alters gene expression within 3 h (Foo et al.,
Sachs and Thimann, 1967; Bangerth, 1994; Li et al., 2005; Fig. 1).
1995; Nordström et al., 2004; Ferguson and Beveridge, In this study, we included parallel observations of
2009). Apically derived auxin has been shown to MAX gene expression and the auxin-responsive gene
down-regulate cytokinin levels in the xylem exudate IAA1. We have also analyzed the axr1-3 mutant, which
of pea and bean (Phaseolus vulgaris), and in pea it can has known defects in auxin-responsive gene expres-
inhibit the expression of ISOPENTYL-TRANSFERASE sion. This allows some discrimination between a direct
cytokinin biosynthesis genes in the stem (Bangerth, effect of strigolactone on MAX3 and MAX4 expression
1994; Li et al., 1995; Tanaka et al., 2006; Ferguson and and an indirect effect mediated by changes in auxin
Beveridge, 2009). In Arabidopsis, the auxin-mediated levels in the mutants. As discussed below, our findings
repression of cytokinin synthesis requires AXR1 suggest that the major portion of feedback regulation
(Nordström et al., 2004). Thus, differences in branch- in Arabidopsis is indirectly mediated by auxin sig-
ing between bdl-2 self-grafts and bdl-2 shoots grafted naling.
Plant Physiol. Vol. 151, 2009 407
Downloaded from www.plantphysiol.org on June 24, 2014 - Published by www.plant.org
Copyright © 2009 American Society of Plant Biologists. All rights reserved.
Hayward et al.

Involvement of Auxin in the Feedback Regulation of MAX3 rootstock) grafts was intermediate between wild-
and MAX4 type/wild-type and max2/max2 grafts (Fig. 5). This
suggests that MAX2 activity in either the shoot or the
Where there were large differences in MAX3 and root can affect MAX3 and MAX4 expression in the
MAX4 transcript accumulation between the wild type rootstock. Although wild-type shoots suppressed
and the max2 mutant, these were substantially reduced MAX3 and MAX4 gene expression in max2 roots
in the axr1-3 mutant background. In contrast, where by about one-third compared with max2 self-grafts,
the differences were small, these appeared to be un- they had no effect on IAA1 gene expression. Thus,
affected by the axr1-3 mutation (Fig. 4). Thus, it seems one possibility is a non-auxin long-distance signaling
likely that a major contributor to the elevated levels of effect. However, the extent of this long-distance regu-
MAX3 and MAX4 transcript in max mutant back- lation is considerably less than that in pea, where
grounds is increased auxin in some tissues, particu- 140-fold differences in RMS1 transcripts were ob-
larly cauline nodes. This is consistent with the very served for the comparable graft combinations (Foo
high DR5:GUS levels observed in the vasculature of et al., 2005). Also unlike comparable grafts of pea, we
max mutant stems and the expression of MAX genes in
vascular-associated cells where polar auxin transport
is occurring (Sorefan et al., 2003; Booker et al., 2005;
Bennett et al., 2006; Stirnberg et al., 2007).
A role for auxin in feedback is further suggested by
the strong correlation between IAA1 expression and
MAX3 and MAX4 expression across the feedback
assays used in this study (Figs. 2, 4, and 5), even in
cases where axr1 had no effect. For example, axr1 did
not suppress the elevated levels of MAX3 and MAX4
transcripts in mature max2 hypocotyls, yet the same
was observed for IAA1 transcripts and additional Aux/
IAAs (Fig. 4B; data not shown). A case for direct or
auxin-independent feedback on MAX3 and MAX4
expression (as in mechanisms 1 or 3 above) is better
suggested where the response of their transcripts to
feedback differs from IAA1. We found only two
such instances. First, in young hypocotyls, the up-
regulation of IAA1 expression in max2 was suppressed
in the axr1-3 background (axr1max2), as expected if
auxin was responsible for its up-regulation, while
MAX3 and MAX4 transcripts were expressed at
equally high levels in max2 and axr1max2 (Fig. 4C).
In this case, however, the enhanced MAX3 and MAX4
expression in axr1max2 relative to axr1 was abolished
by NPA treatment (Fig. 4D). This could again mean a
function for increased auxin in the polar auxin trans-
port stream (PATS) of max2 plants in this feedback
(Bennett et al., 2006) or that transcript levels are at their
lowest after NPA treatment, preventing an additional
effect of MAX2. The other case where IAA1 expression
levels did not reflect MAX3 and MAX4 expression was
in the rootstock of grafted plants, as discussed below Figure 6. Interactions between auxin and the MAX pathway. Regula-
(Fig. 5B). tory signals represented in blue were investigated in this study. Step 1,
Auxin traveling in the PATS promotes MAX3 and MAX4 expression in
an AXR1-dependent manner, prevented if IAA12 is stabilized. Step 2,
Auxin-promoted MAX3 and MAX4 transcript levels lead to increased
Systemic Feedback Regulation of MAX3 and MAX4
strigolactone production, which leads to reduced bud outgrowth and
The data described above demonstrate that auxin, reduced auxin export from buds into the PATS. Step 3, If strigolactones
likely that traveling in the PATS, is a major component are low, branching is increased and auxin level increases, feedback up-
regulating MAX3 and MAX4 to increase strigolactone levels as in step
of feedback signaling on MAX3 and MAX4 transcript
1. Step 4, Strigolactone signaling may feedback down-regulate tran-
abundance, especially in cauline nodes. Therefore, at script levels independent of auxin; the dashed line indicates weak/
least in the shoot, auxin could act as a long-range putative interaction. Step 5, Auxin also down-regulates cytokinin
feedback signal for MAX action. synthesis, and cytokinin promotes branching. The environment/geno-
In this study, similar to those for pea (Foo et al., 2005; type/developmental program may influence each of these regulatory
Johnson et al., 2006), the expression of MAX3 and stages tissue specifically. [See online article for color version of this
MAX4 in the rootstocks of wild-type/max2 (scion/ figure.]

408 Plant Physiol. Vol. 151, 2009


Downloaded from www.plantphysiol.org on June 24, 2014 - Published by www.plant.org
Copyright © 2009 American Society of Plant Biologists. All rights reserved.
Auxin and Strigolactone Interactions in Branching

did not detect a systemic feedback signal from max2 2007). Although many similarities are apparent be-
shoots that could up-regulate MAX3 and MAX4 ex- tween species, differences in strigolactone pathways
pression in wild-type rootstocks (Fig. 5; Foo et al., could potentially reflect differences in developmental
2005). Interestingly, in petunia, the up-regulation of the programs and shoot architectures.
MAX4 ortholog, DECREASED APICAL DOMINANCE1,
was not found to be graft transmissible to the rootstock
(Simons et al., 2007).
The relatively minor effect of the shoot genotype, CONCLUSION
wild type or max2, on MAX3 and MAX4 gene expres-
sion in the roots contrasts with clear evidence of We demonstrate that the auxin regulation of strigo-
the long-distance regulation of xylem sap cytokinin lactone biosynthetic genes is conserved in Arabidop-
by strigolactone signaling in Arabidopsis shoots sis. This involves AXR1/TIR1-regulated Aux/IAA
(Foo et al., 2007). Auxin regulation of cytokinin content stability and contributes to branching inhibition. Fur-
is well documented (Nordström et al., 2004), and thermore, this auxin regulation also acts as a feedback
decapitation leads to depletion of xylem sap cytoki- mechanism, with increased auxin content in condi-
nins, which can be restored with exogenous auxin tions of low strigolactone acting as a downstream
(Bangerth, 1994; Li et al., 1995). However, a feedback communicator to increase strigolactone biosynthesis.
signal in addition to auxin has been proposed for pea, We suggest that auxin-regulated strigolactone biosyn-
because feedback down-regulated xylem cytokinin is thesis may form a conserved component of auxin-
not significantly increased by decapitation in the rms1, mediated branching inhibition and that auxin and
rms3, and rms4 mutants over 24 h, and the up to 2,000- strigolactone signaling may participate in an inter-
fold increases in RMS1 expression observed in the locking feedback loop, involving interplay with addi-
rms4 (Psmax2) mutant have not been achieved by tional stimuli, to precisely control branching in plants.
auxin addition (Foo et al., 2005, 2007; for review, see A diagrammatic representation of the interactions
Dun et al., 2009). More direct studies of auxin involve- between auxin and the MAX pathway built upon by
ment in shoot-to-root regulation of xylem sap cytoki- this study is shown in Figure 6.
nin levels in Arabidopsis are required to determine
whether a signal in addition to auxin affects xylem
cytokinin content.
MATERIALS AND METHODS

Sources of Differences Plant Growth and Materials


For Figures 1, 2, 3E to 3G, 4C, 4D, and 5, Arabidopsis (Arabidopsis thaliana)
In this and previous studies, similar experiments
seeds were sown on a mixture of California potting mix type C and vermic-
have given quantitatively and sometimes qualitatively ulite (3:2, v/v) at a density of one to two per 5 cm2 or one per 1 cm2 (Fig. 2) and
different results. Our data suggest that some of the stratified for 2 to 3 d at 4°C. Trays were transferred either to a temperature-
variability is likely technical, such as the method of controlled growth room at 22°C 6 2°C/18°C 6 2°C for 18 h of light/6 h of dark
hormone application or expression analysis, the hor- with fluorescent lighting (Philips 36 W/840) supplying approximately 120
mmol m22 s21 light or to a temperature-controlled glasshouse at 24°C 6 2°C/
mone concentrations used, and the tissue types in- 18°C 6 2°C for 18 h of light/6 h of dark with the natural daylength extended
cluded in the analyses. Different phenotypes of species by incandescent light. For Figures 3A to 3D, 4A, and 4B, seeds were cold
dictate the use of different tissues for grafting and gene treated and grown at a density of one per 4 cm2 in a temperature-controlled
expression analysis, epicotyl in pea versus hypocotyl glasshouse according to Bennett et al. (2006). All plant lines used were in the
Col-0 background as follows: Col-0 wild type, max1-1, max2-1, max3-11, max4-1,
in Arabidopsis and petunia, and perhaps this could axr1-3, axr1-3max2-1, and bdl-2 (see below). Vegetative 2-week-old plants had
influence the movement of, or degree of response to, approximately six to eight mature leaves expanded, and bolting plants were
long-distance signals. Importantly, some interesting grown until the primary bolt was 2 to 5 cm (3 weeks) or 15 to 30 cm
biological differences are also emerging. In pea but not (approximately 5 weeks) as indicated in the text.
Arabidopsis, mutant rms4 (max2) rootstocks are more
able than wild-type rootstocks to suppress branching The bdl-2 Allele
in wild-type and various rms mutant shoots, and this
The bdl-2 mutant allele was found in a screen for branching mutants of
has been associated with the small local feedback up- the ethyl methanesulfonate-mutagenized Col-0 population described by
regulation of strigolactone biosynthesis genes in mu- Stirnberg et al. (2002). Genetic analysis of the mutant individual recovered
tant rootstocks (Beveridge et al., 1996, 1997b; Morris showed that its max-like phenotype was caused by a semidominant mutation
et al., 2001; Booker et al., 2005; Foo et al., 2005; Johnson in a heterozygous state. Homozygotes were severely dwarfed. The mutation
mapped between PVV4 and nga63 on chromosome I, an interval containing
et al., 2006). While this could involve differences in the
four IAA genes: IAA3/SHY2, IAA17/AXR3, IAA10, and IAA12/BDL. As muta-
strength of feedback regulation, perhaps strigolactone tions with semidominant inheritance were known for this class of genes but
levels could be more rate limiting in pea than in homozygous or heterozygous shy2 or axr3 were unlike our mutant (Leyser
Arabidopsis in standard conditions. As discussed et al., 1996; Tian and Reed, 1999), IAA10 and IAA12 were considered as
above, the balance of auxin and auxin-independent possible candidates and their coding regions were amplified from genomic
DNA of homozygous mutant plants. Sequencing revealed no change in IAA10
feedback regulation on strigolactone synthesis genes but a CCA-to-TCA change in codon 74 of IAA12, predicting a Pro-to-Ser
may also differ, the latter component appearing to be substitution identical to the bdl-1 allele, which is in the Landsberg erecta
greater in pea than in Arabidopsis (Foo et al., 2005, genetic background (Hamann et al., 1999, 2002).

Plant Physiol. Vol. 151, 2009 409


Downloaded from www.plantphysiol.org on June 24, 2014 - Published by www.plant.org
Copyright © 2009 American Society of Plant Biologists. All rights reserved.
Hayward et al.

Decapitation and Hormone Treatments was performed using Power SYBR Green PCR Master Mix in an ABI
Prism 7300 Sequence Detection System (Applied Biosystems) with melt-curve
Three-week-old plants were decapitated by removing the entire primary analysis. Gene expression was normalized to 18S using the comparative
inflorescence (2–5 cm) and the youngest three to four expanding rosette leaves CT method with the formula 22CTgene/22CT18S. Primer sequences are as fol-
using a pair of fine forceps. For 5-week-old plants, primary bolts (15–30 cm) lows: 18SF (5#-TTCCTAGTAAGCGCGAGTCATCA-3#), 18SR (5#-GAACA-
were decapitated 15 mm above the stem base using a sterile scalpel, with or CTTCACCGGATCATTCAAT-3#), QRTMAX4F (5#-GAAAGATACCCACT-
without the youngest rosette leaves removed as indicated in the figure TGGCTGAATG-3#) QRTMAX4R (5#-TGTGGAGTAGCCGTCGAAGAG-3#),
legends. For hormone treatments, hormones were dissolved in ethanol and QRTMAX3F (5#-GATTCGTTGGTGAGCCCATG-3#), QRTMAX3R (5#-CAC-
mixed with lanolin to give the various concentrations indicated with a final CGAAACCGCATACTCGA-3#), QRTIAA1F (5#-GCTCCTCCTCCTGCAAA-
ethanol concentration of less than 10%. NPA was applied in a ring either AACAC-3#), QRTIAA1R (5#-ACGGTTAGATCTCACTGGAGGC-3#); b-actin
around the top of the hypocotyls or around the primary bolt 15 mm above the primers were as described by Brown et al. (2003). The Genevestigator V3
base as indicated. IAA was applied to the decapitation site or, for the dose- Classic Metaprofile analysis tool (available at https://www.genevestigator.
response assay, in a ring around the primary bolt of intact plants 15 mm above ethz.ch/) was used to generate an anatomy heat map of IAA1 (245397_at),
the base. Control treatments were as above minus hormone. For expression MAX3 (266129_at), MAX4 (253398_at), and IAA12 (264605_at; data not
analyses, tissues as indicated (approximately 2–5 mm of hypocotyl, the basal shown).
10 mm of the primary inflorescence stem, and the first 5 cm of root tissue) were
harvested after the times indicated. Statistical P values for all phenotypic and
expression data were calculated using the t test.
Plants were treated with GR24 (http://www.chiralix.com/) as described ACKNOWLEDGMENTS
by Gomez-Roldan et al. (2008) with modifications. For 5 mM-treated and 0 mM
At the University of Queensland, we thank Julia Cremer and the
control plants, solutions containing 0.05% Silwet and 0.05% acetone with or
Beveridge laboratory for assistance with the experiment in Figure 2, Dr.
without GR24 were applied to plants every 3 d starting at 18 d after transfer to
Philip Brewer for assistance with the strigolactone application experiment,
the growth room. At 24 d, the GR24 dose was doubled to 10 mM applied every
Dr. Elizabeth Dun and Tanya Brcich for helpful discussions, Mr. Bob Simpson
2 d for two treatments followed by an additional four treatments of 5 mM every
for qRT-PCR advice, and Kerry Condon for technical assistance. At the
3 d. Rosette leaf numbers (minus cotyledons) were scored 1 week after bolting,
University of York, we thank horticultural staff for plant care, Dr. Karin van
and branch numbers of 5 mm or greater were scored at 47 to 50 d.
de Sande for initial isolation of the bdl-2 mutant, and University of York
Technology Facility staff for qRT-PCR advice. We are also indebted to Dr.
Grafting Catherine Rameau (INRA, France), who provided the GR24.
Received February 24, 2009; accepted July 21, 2009; published July 29, 2009.
Transverse grafts were performed using a protocol adapted from Turnbull
et al. (2002). Sterilized cold-treated seeds were germinated vertically on plates
containing half-strength Murashige and Skoog basal salts, 1% Suc, and 0.8%
agar in an incubator at 24°C with an 18-h photoperiod. Seedlings were grown LITERATURE CITED
for 4 to 5 d before careful transfer to sterilized petri dishes containing a
premoistened layer of Millipore nitrocellulose filter (type HA; pore, 0.45 mm) Abel S, Nguyen MD, Theologis A (1995) The PS-IAA4/5-like family of
on a single layer of Whatman No. 1 filter paper for grafting. Grafts were early auxin-inducible mRNAs in Arabidopsis thaliana. J Mol Biol 251:
returned to the incubator, and successful grafts were transferred to soil by 7 d 533–549
and grown in a temperature-controlled growth room as above. Grafts were Akiyama K, Matsuzaki K, Hayashi H (2005) Plant sesquiterpenes induce
visually assessed, and phenotypic analyses were carried out at the cessation of hyphal branching in arbuscular mycorrhizal fungi. Nature 435: 824–827
primary meristem activity. For expression analyses, hypocotyl and root tissues Arite T, Iwata H, Ohshima K, Maekawa M, Nakajima M, Kojima M,
below the graft union (approximately 2 cm) were collected as rootstock tissue Sakakibara H, Kyozuka J (2007) DWARF10, an RMS1/MAX4/DAD1
and rosette stem tissue (rosette stem plus approximately 0.5 mm of petiole ortholog, controls lateral bud outgrowth in rice. Plant J 51: 1019–1029
tissue) was collected as shoot tissue. For vegetative-stage plants, tissues were Auldridge ME, Block A, Vogel JT, Dabney-Smith C, Mila I, Bouzayen M,
collected 2 weeks after transfer to soil. For mature plants, tissues were Magallanes-Lundback M, DellaPenna D, McCarty DR, Klee HJ (2006)
collected following phenotypic analysis. Characterisation of three members of the Arabidopsis carotenoid cleav-
age dioxygenase family demonstrates the divergent roles of this mul-
tifunctional enzyme family. Plant J 45: 982–993
qRT-PCR Bainbridge K, Sorefan K, Ward S, Leyser O (2005) Hormonally controlled
expression of the Arabidopsis MAX4 shoot branching regulatory gene.
Total RNA was isolated from pools of five to 25 plants using NucleoSpin Plant J 44: 569–580
RNA plant kits (Machery-Nagel) and quantified using a NanoDrop 1000. Bangerth F (1994) Response of cytokinin concentration in the xylem
cDNA was synthesized in 20 mL with 100 ng to 1 mg of total RNA, 250 ng of exudate of bean (Phaseolus vulgaris L.) plants to decapitation and auxin
random primers (Promega), 250 ng of oligo(dT)15 (Promega), 0.5 mM deoxy- treatment and relationship to apical dominance. Planta 194: 439–442
ribonucleotide triphosphates, 5 mM dithiothreitol (Invitrogen), 13 first-strand Bennett T, Sieberer T, Willett B, Booker J, Luschnig C, Leyser O (2006) The
buffer, and 100 units of SuperScript III reverse transcriptase (Invitrogen) as Arabidopsis MAX pathway controls shoot branching by regulating
described in the Invitrogen SSIII RT protocol. Reverse transcriptase-minus auxin transport. Curr Opin Plant Biol 16: 553–563
control reactions were performed for each RNA sample to assess DNA Beveridge CA (2000) Long-distance signalling and mutational analysis of
contamination during qRT-PCR. Each qRT-PCR was performed in duplicate branching in pea. Plant Growth Regul 32: 193–203
using SYBR Green PCR Master Mix, 200 nM of each primer, and 2 to 15 ng of Beveridge CA, Murfet IC, Kerhoas L, Sotta B, Miginiac E, Rameau C
cDNA in an ABI Prism 7900HT Sequence Detection System (Applied Biosys- (1997a) The shoot controls zeatin riboside export from pea roots:
tems) with melt-curve analysis. No-template control reactions were per- evidence from the branching mutant rms4. Plant J 11: 339–345
formed for each primer. Primer efficiencies (PE) of each gene were calculated Beveridge CA, Ross JJ, Murfet IC (1996) Branching in pea: action of genes
per run by LinRegPCR (Ramakers et al., 2003). Relative quantitation was Rms3 and Rms4. Plant Physiol 110: 859–865
determined using the comparative cycle threshold (CT) method; CT values for Beveridge CA, Symons GM, Murfet IC, Ross JJ, Rameau C (1997b) The
technical replicates for each gene were averaged (AvCT), and MAX3 rms1 mutant of pea has elevated indole-3-acetic acid levels and reduced
(At2g44990), MAX4 (At4g32810), and IAA1 (At4g14560) expression was root-sap zeatin riboside content but increased branching controlled by
normalized against 18S (At2g01010; Figs. 1, 2, B–D, and 3–5) or three b-actin graft-transmissible signals. Plant Physiol 115: 1251–1258
reference genes (At3g18780, At5g09810, and At1g49240; Fig. 2A) according to Booker J, Auldridge M, Wills S, McCarty D, Klee H, Leyser O (2004)
Brown et al. (2003) using the formula PEgene2AvCTgene/PEreference2AvCTreference. MAX3/CCD7 is a carotenoid cleavage dioxygenase required for the
Resulting values were averaged for biological replicates and plotted with SE synthesis of a novel plant signalling molecule. Curr Biol 14: 1232–1238
values relative to a chosen sample (the calibrator), as indicated on the y axis. Booker J, Sieberer T, Wright W, Williamson L, Willett B, Stirnberg P,
Primers were designed using PrimerExpress 1.5 (Applied Biosystems) with Turnbull C, Srinivasan M, Goddard P, Leyser O (2005) MAX1 encodes a
one primer across intron junctions where possible. The qRT-PCR for Figure 3 cytochrome P450 family member that acts downstream of MAX3/4 to

410 Plant Physiol. Vol. 151, 2009


Downloaded from www.plantphysiol.org on June 24, 2014 - Published by www.plant.org
Copyright © 2009 American Society of Plant Biologists. All rights reserved.
Auxin and Strigolactone Interactions in Branching

produce a carotenoid-derived branch-inhibiting hormone. Dev Cell 8: Humphrey AJ, Beale MH (2006) Strigol: biogenesis and physiological
443–449 activity. Phytochemistry 67: 636–640
Brewer P, Dun E, Ferguson B, Rameau C, Beveridge C (2009) Strigolactone Ishikawa S, Maekawa M, Arite T, Onishi K, Takamure I, Kyozuka J (2005)
acts downstream of auxin to regulate bud outgrowth in pea and Suppression of tiller bud activity in tillering dwarf mutants of rice. Plant
Arabidopsis. Plant Physiol 150: 482–493 Cell Physiol 46: 79–86
Brown RL, Kazan K, McGrath KC, Maclean DJ, Manners JM (2003) A role Johnson AW, Gowda G, Hassanali A, Knox J, Monaco S, Razavi Z,
for the GCC-box in jasmonate-mediated activation of the PDF1.2 gene of Rosebery G (1981) The preparation of synthetic analogs of strigol.
Arabidopsis. Plant Physiol 132: 1020–1032 J Chem Soc Perkin Trans 1 1981: 1734–1743
Catterou M, Dubois F, Smets R, Vaniet S, Kichey T, Van Onckelen H, Johnson X, Brcich T, Dun EA, Goussot M, Haurogné K, Beveridge CA,
Sangwan-Norreel BS, Sangwan RS (2002) hoc: an Arabidopsis mutant Rameau C (2006) Branching genes are conserved across species: Genes
overproducing cytokinins and expressing high in vitro organogenic controlling a novel signal in pea are coregulated by other long-distance
capacity. Plant J 30: 273–287 signals. Plant Physiol 142: 1014–1026
Chaudhury AM, Letham S, Dennis ES (1993) amp1: a mutant with high Kepinski S, Leyser O (2005) The Arabidopsis F-box protein TIR1 is an
cytokinin levels and altered embryonic pattern, faster vegetative auxin receptor. Nature 435: 446–451
growth, constitutive photomorphogenesis and precocious flowering. Lazar G, Goodman HM (2006) MAX1, a regulator of the flavonoid pathway,
Plant J 4: 907–916 controls vegetative axillary bud outgrowth in Arabidopsis. Proc Natl
Cook CE, Whichard LP, Wall ME, Egley GH, Coggon P, Luhan PA, Acad Sci USA 103: 472–476
McPhail AT (1972) Germination stimulants. II. The structure of strigol: a Leyser HM, Lincoln CA, Timpte C, Lammer D, Turner J, Estelle M (1993)
potent seed germination stimulant for witchweed (Striga lutea Lour.). Arabidopsis auxin-resistance gene AXR1 encodes a protein related to
J Am Chem Soc 94: 6198–6199 ubiquitin-activating enzyme E1. Nature 364: 161–164
del Pozo JC, Dharmasiri S, Hellmann H, Walker L, Gray WM, Estelle M Leyser HMO, Pickett FB, Dharmasiri S, Estelle M (1996) Mutations in the
(2002) AXR1-ECR1-dependent conjugation of RUB1 to the Arabidopsis AXR3 gene of Arabidopsis result in altered auxin response including
cullin AtCUL1 is required for auxin response. Plant Cell 14: 421–433 ectopic expression from the SAUR-AC1 promoter. Plant J 10: 403–413
Dharmasiri N, Dharmasiri S, Estelle M (2005a) The F-box protein TIR1 is Li CJ, Herrera GJ, Bangerth F (1995) Effect of apex excision and replace-
an auxin receptor. Nature 435: 441–445 ment by 1-naphthylacetic acid on cytokinin concentration and apical
Dharmasiri N, Dharmasiri S, Weijers D, Karunarathna N, Jurgens G, dominance in pea plants. Physiol Plant 94: 465–469
Estelle M (2007) AXL and AXR1 have redundant functions in RUB Lincoln C, Britton JH, Estelle M (1990) Growth and development of the
conjugation and growth and development in Arabidopsis. Plant J 52: axrl mutants of Arabidopsis. Plant Cell 2: 1071–1080
114–123 Ljung K, Bhalerao RP, Sandberg G (2001) Sites and homeostatic control of
Dharmasiri N, Dharmasiri S, Weijers D, Lechner E, Yamada M, Hobbie L, auxin biosynthesis in Arabidopsis during vegetative growth. Plant J 28:
Ehrismann JS, Jürgens G, Estelle M (2005b) Plant development is regu- 465–474
lated by a family of auxin receptor F box proteins. Dev Cell 9: 109–119 McKenzie MJ, Mett VV, Stewart Reynolds PH, Jameson PE (1998) Con-
Dharmasiri S, Estelle M (2002) The role of regulated protein degradation in trolled cytokinin production in transgenic tobacco using a copper-
auxin response. Plant Mol Biol 9: 401–409 inducible promoter. Plant Physiol 116: 969–977
Dun EA, Brewer PB, Beveridge CA (2009) Strigolactones: discovery of the Morris SE, Turnbull CGN, Murfet IC, Beveridge CA (2001) Mutational
elusive shoot branching hormone. Trends Plant Sci 14: 364–372 analysis of branching in pea: evidence that Rms1 and Rms5 regulate the
Dun EA, Ferguson BJ, Beveridge CA (2006) Apical dominance and shoot same novel signal. Plant Physiol 126: 1205–1213
branching: divergent opinions or divergent mechanisms. Plant Physiol Nordström A, Tarkowski P, Tarkowska D, Norbaek R, Åstot C, Dolezal K,
142: 812–819 Sandberg G (2004) Auxin regulation of cytokinin biosynthesis in Arab-
Faiss M, Zalubilova J, Strnad M, Schmülling T (1997) Conditional trans- idopsis thaliana: a factor of potential importance for auxin-cytokinin-
genic expression of the ipt gene indicates a function for cytokinins in regulated development. Proc Natl Acad Sci USA 101: 8039–8044
paracrine signalling in whole tobacco plants. Plant J 12: 401–415 Ongaro V, Leyser O (2007) Hormonal control of shoot branching. J Exp Bot
Ferguson BJ, Beveridge CA (2009) Roles for auxin, cytokinin and strigo- 59: 67–74
lactone in regulating shoot branching. Plant Physiol 149: 1929–1944 Park JY, Kim HJ, Kim J (2002) Mutation in domain II of IAA1 confers
Foo E, Bullier E, Goussot M, Foucher F, Rameau C, Beveridge CA (2005) diverse auxin-related phenotypes and represses auxin-activated expres-
The branching gene RAMOSUS1 mediates interactions among two sion of Aux/IAA genes in steroid regulator-inducible system. Plant J 32:
novel signals and auxin in pea. Plant Cell 17: 464–474 669–683
Foo E, Morris SE, Parmenter K, Young N, Wang H, Jones A, Rameau C, Petroski MD, Deshaies RJ (2005) Function and regulation of cullin-ring
Turnbull CGN, Beveridge CA (2007) Feedback regulation of xylem ubiquitin ligases. Nat Rev Mol Cell Biol 6: 9–20
cytokinin content is conserved in pea and Arabidopsis. Plant Physiol Ramakers C, Ruijter JM, Lekanne Deprez LH, Moorman AFM (2003)
143: 1418–1428 Assumption-free analysis of quantitative real-time PCR data. Neurosci
Gomez-Roldan V, Fermas S, Brewer PB, Puech-Pagès V, Dun EA, Pillot JP, Lett 339: 62–66
Letisse F, Matusova R, Danoun S, Portais JC, et al (2008) Strigolactone Romano CP, Robson PR, Smith H, Estelle M, Klee H (1995) Transgene-
inhibition of shoot branching. Nature 455: 189–194 mediated auxin overproduction in Arabidopsis: hypocotyl elongation
Gray WM, Kepinski S, Rouse D, Leyser O, Estelle M (2001) Auxin phenotype and interactions with the hy6-1 hypocotyl elongation and
regulates SCFTIR1-dependent degradation of AUX/IAA proteins. Na- axr1 auxin-resistant mutants. Plant Mol Biol 27: 1071–1083
ture 414: 271–276 Sachs T, Thimann K (1967) The role of auxins and cytokinins in the release
Hamann T, Benkova E, Baurle I, Kientz M, Juergens G (2002) The of buds from dominance. Am J Bot 54: 136–144
Arabidopsis BODENLOS gene encodes an auxin response protein Sakamoto T, Sakakibara H, Kojima M, Yamamoto Y, Nagasaki H, Inukai
inhibiting MONOPTEROS-mediated embryo patterning. Genes Dev Y, Sato Y, Matsuoka M (2006) Ectopic expression of KNOTTED1-like
16: 1610–1615 homeobox protein induces expression of cytokinin biosynthesis genes in
Hamann T, Mayer U, Jurgens G (1999) The auxin-insensitive bodenlos rice. Plant Physiol 142: 54–62
mutation affects primary root formation and apical-basal patterning in Schmitz G, Theres K (2005) Shoot and inflorescence branching. Curr Opin
the Arabidopsis embryo. Development 126: 1387–1395 Plant Biol 8: 506–511
Hardtke CS, Ckurshumova W, Vidaurre DP, Singh SA, Stamatiou G, Schmülling T (2002) New insights into the functions of cytokinins in plant
Tiwari SB, Hagen G, Guilfoyle TJ, Berleth T (2004) Overlapping and development. J Plant Growth Regul 21: 40–49
non-redundant functions of the Arabidopsis auxin response factors Schwechheimer C, Serino G, Callis J, Crosby WL, Lyapina S, Deshaies RJ,
MONOPTEROS and NONPHOTOTROPIC HYPOCOTYL 4. Develop- Gray WM, Estelle M, Deng XW (2001) Interactions of the COP9
ment 131: 1089–1100 signalosome with the E3 ubiquitin ligase SCFTIR1 in mediating auxin
Hruz T, Laule O, Szabo G, Wessendrop F, Bleuler S, Oertle L, Widmayer P, response. Science 292: 1379–1382
Gruissem W, Zimmermann P (2008) Genevestigator V3: a reference Schwechheimer C, Serino G, Deng X (2002) Multiple ubiquitin-mediated
expression database for the meta-analysis of transcriptomes. Advances processes require COP9 signalosome and AXR1 function. Plant Cell 14:
in Bioinformatics 2008: 420747 2553–2563

Plant Physiol. Vol. 151, 2009 411


Downloaded from www.plantphysiol.org on June 24, 2014 - Published by www.plant.org
Copyright © 2009 American Society of Plant Biologists. All rights reserved.
Hayward et al.

Simons JL, Napoli CA, Janssen BJ, Plummer KM, Snowden KC (2007) Turnbull CG, Booker JP, Leyser O (2002) Micrografting techniques for
Analysis of the DECREASED APICAL DOMINANCE genes of petunia testing long-distance signalling in Arabidopsis. Plant J 32: 255–262
in the control of axillary branching. Plant Physiol 143: 697–706 Ulmasov T, Hagen G, Guilfoyle TJ (1999) Activation and repression of
Snowden KC, Simkin AJ, Janssen BJ, Templeton KR, Loucas HM, transcription by auxin-response factors. Proc Natl Acad Sci USA 96:
Simons JL, Karunairetnam S, Gleave AP, Clark DG, Klee HJ (2005) 5844–5849
The Dad1/PhCCD8 gene affects branch production and has a role in Umehara M, Hanada A, Yoshida S, Akiyama K, Arite T, Takeda-Kamiya
leaf senescence, root growth and flower development. Plant Cell 17: N, Magome H, Kamiya Y, Shirasu K, Yoneyama K, et al (2008)
746–759 Inhibition of shoot branching by new terpenoid plant hormones. Nature
Sorefan K, Booker J, Haurogne K, Goussot M, Bainbridge K, Foo E, 455: 195–200
Chatfield SP, Ward S, Beveridge CA, Rameau C, et al (2003) MAX4 and Wang KLC, Li H, Ecker JR (2002) Ethylene biosynthesis and signalling
RMS1 are orthologous dioxygenase-like genes that regulate shoot networks. Plant Cell (Suppl) 14: S131–S151
branching in Arabidopsis and pea. Genes Dev 17: 1469–1474 Wickson ME, Thimann KV (1958) The antagonism of auxin and kinetin in
Stirnberg P, Furner IJ, Leyser HMO (2007) MAX2 participates in an SCF apical dominance. Physiol Plant 11: 62–74
complex which acts locally at the node to suppress shoot branching. Wu K, Chen A, Pan ZQ (2000) Conjugation of Nedd8 to CUL1 enhances the
Plant J 50: 80–94 ability of the ROC1-CUL1 complex to promote ubiquitin polymeriza-
Stirnberg P, van de Sande K, Leyser HMO (2002) MAX1 and MAX2 control tion. J Biol Chem 275: 32317–32324
shoot lateral branching in Arabidopsis. Development 129: 1131–1141 Yamaguchi S (2008) Gibberellin metabolism and its regulation. Annu Rev
Tan X, Calderon-Villalobos LI, Sharon M, Zheng C, Robinson CV, Estelle Plant Biol 59: 225–251
M, Zheng N (2007) Mechanism of auxin perception by the TIR1 ubiq- Yang X, Lee S, So JH, Dharmasiri S, Dharmasiri N, Ge L, Jensen C,
uitin ligase. Nature 446: 640–645 Hangarter R, Hobbie L, Estelle M (2004) The IAA1 protein is encoded
Tanaka M, Takei K, Kojima M, Sakakibara H, Mori H (2006) Auxin by AXR5 and is a substrate of SCFTIR1. Plant J 40: 772–782
controls local cytokinin biosynthesis in the nodal stem in apical dom- Zenser N, Ellsmore A, Leasure C, Callis J (2001) Auxin modulates the
inance. Plant J 45: 1028–1036 degradation rate of AUX/IAA proteins. Proc Natl Acad Sci USA 98:
Thimann KV, Skoog F (1934) On the inhibition of bud development and 11795–11800
other functions of growth substance in Vicia faba. Proc R Soc Lond B Biol Zimmermann P, Hirsch-Hoffmann M, Hennig L, Gruissem W (2004)
Sci 114: 317–339 GENEVESTIGATOR: Arabidopsis microarray database and analysis
Tian Q, Reed JW (1999) Control of auxin-regulated root development by toolbox. Plant Physiol 136: 2621–2632
the Arabidopsis thaliana SHY2/IAA3 gene. Development 126: 711–721 Zou J, Chen Z, Zhang S, Zhang W, Jiang G, Zhao X, Zhai W (2005)
Tiryaki I, Staswick PE (2002) An Arabidopsis mutant defective in jasm- Characterizations and fine mapping of a mutant gene for high tillering
onate response is allelic to the auxin-signaling mutant axr1. Plant and dwarf in rice (Oryza sativa L.). Planta 222: 604–612
Physiol 130: 887–894 Zou J, Zhang S, Zhang W, Li G, Chen Z, Zhai W, Zhao X, Pan X, Xie Q, Zhu
Tiwari SB, Wang XJ, Hagen G, Guilfoyle TJ (2001) AUX/IAA proteins are L (2006) The rice HIGH-TILLERING DWARF1 encoding an orthologue
active repressors, and their stability and activity are modulated by of Arabidopsis MAX3 is required for negative regulation of the out-
auxin. Plant Cell 13: 2809–2822 growth of axillary buds. Plant J 48: 687–696

412 Plant Physiol. Vol. 151, 2009


Downloaded from www.plantphysiol.org on June 24, 2014 - Published by www.plant.org
Copyright © 2009 American Society of Plant Biologists. All rights reserved.

You might also like