03.12.2012 Views

recent discoveries suggest distinct pathways and novel mechanisms

recent discoveries suggest distinct pathways and novel mechanisms

recent discoveries suggest distinct pathways and novel mechanisms

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

CSIRO PUBLISHING<br />

www.publish.csiro.au/journals/fpb Functional Plant Biology, 2007, 34, 465–473<br />

Review:<br />

Starch breakdown: <strong>recent</strong> <strong>discoveries</strong> <strong>suggest</strong> <strong>distinct</strong> <strong>pathways</strong><br />

<strong>and</strong> <strong>novel</strong> <strong>mechanisms</strong><br />

Samuel C. Zeeman A,B , Thierry Delatte A ,Gaëlle Messerli A , Martin Umhang A , Michaela Stettler A ,<br />

Tabea Mettler A , Sebastian Streb A , Heike Reinhold A <strong>and</strong> Oliver Kötting A<br />

A Institute of Plant Sciences, ETH Zurich, Universitätstrasse 2, CH-8092 Zurich, Switzerl<strong>and</strong>.<br />

B Corresponding author. Email: szeeman@ethz.ch<br />

This paper originates from a presentation at the 8th International Congress of Plant<br />

Molecular Biology, Adelaide, Australia, August 2006.<br />

Abstract. The aim of this article is to provide an overview of current models of starch breakdown in leaves. We summarise<br />

the results of our <strong>recent</strong> work focusing on Arabidopsis, relating them to other work in the field. Early biochemical studies<br />

of starch containing tissues identified numerous enzymes capable of participating in starch degradation. In the non-living<br />

endosperms of germinated cereal seeds, starch breakdown proceeds by the combined actions of α-amylase, limit dextrinase<br />

(debranching enzyme), β-amylase <strong>and</strong> α-glucosidase. The activities of these enzymes <strong>and</strong> the regulation of some of the<br />

respective genes on germination have been extensively studied. In living plant cells, additional enzymes are present, such<br />

as α-glucan phosphorylase <strong>and</strong> disproportionating enzyme, <strong>and</strong> the major pathway of starch breakdown appears to differ<br />

from that in the cereal endosperm in some important aspects. For example, reverse-genetic studies of Arabidopsis show<br />

that α-amylase <strong>and</strong> limit-dextrinase play minor roles <strong>and</strong> are dispensable for starch breakdown in leaves. Current data<br />

also casts doubt on the involvement of α-glucosidase. In contrast, several lines of evidence point towards a major role<br />

for β-amylase in leaves, which functions together with disproportionating enzyme <strong>and</strong> isoamylase (debranching enzyme)<br />

to produce maltose <strong>and</strong> glucose. Furthermore, the characterisation of Arabidopsis mutants with elevated leaf starch has<br />

contributed to the discovery of previously unknown proteins <strong>and</strong> metabolic steps in the pathway. In particular, it is now<br />

apparent that glucan phosphorylation is required for normal rates of starch mobilisation to occur, although a detailed<br />

underst<strong>and</strong>ing of this step is still lacking. We use this review to give a background to some of the classical genetic mutants<br />

that have contributed to our current knowledge.<br />

Additional keywords: amylase, amylopectin, Arabidopsis thaliana L., carbohydrate metabolism, cereal endosperm,<br />

metabolic regulation.<br />

Starch is central to Arabidopsis metabolism<br />

Starch <strong>and</strong> sucrose are the primary products of photosynthesis<br />

in the leaves of most plants. Triose-phosphates exported from<br />

the chloroplast during the day are used to synthesise sucrose<br />

in the cytosol. Sucrose is then loaded into the phloem of the<br />

source tissues (e.g. mature leaves) for long-distance transport to<br />

sink tissues (e.g. roots, developing leaves <strong>and</strong> storage organs)<br />

where it is used to support respiration <strong>and</strong> growth. Starch<br />

is synthesised concurrently with sucrose in the leaves. It is<br />

made in the chloroplast <strong>and</strong> comprises glucose polymers that<br />

are packed to form dense, osmotically inert granules (Buléon<br />

et al. 1998; Zeeman et al. 2002). Arabidopsis plants grown in<br />

st<strong>and</strong>ard growth room conditions (e.g. a 12-h photoperiod with<br />

150 µmoles photon m−2 s−1 quantum irradiance, 20◦C, 70%<br />

relative humidity) partition ∼40% of their newly-assimilated<br />

carbon into starch (Zeeman <strong>and</strong> ap Rees 1999). Leaf starch is<br />

degraded throughout the night. The sugars released are used to<br />

provide substrates for respiration of the leaf <strong>and</strong> for continued<br />

sucrose synthesis <strong>and</strong> export (Zeeman <strong>and</strong> ap Rees 1999).<br />

The division of assimilates between starch <strong>and</strong> sucrose is<br />

controlled in accordance with the environmental conditions.<br />

Chatterton <strong>and</strong> Silvius (1980) showed that several plant species<br />

grown under short-day conditions partition relatively more of<br />

their photo-assimilates into starch than those grown under longday<br />

conditions (presumably at the expense of daytime sucrose<br />

production). This has also been shown to be the case for<br />

Arabidopsis (Gibon et al. 2004). Thus, although the total amount<br />

of carbon assimilated during the short day is lower than in a<br />

long day, the amount of starch accumulated during the entire<br />

photoperiod does not decrease to the same extent (Fig. 1A).<br />

A greater proportion of the photo-assimilated carbon is therefore<br />

available to sustain metabolism during the long night that follows<br />

a short day. During the night, starch is broken down at a relatively<br />

constant rate. The rate of breakdown is regulated in a manner<br />

complementary to the rate of synthesis during the day. When<br />

plants are entrained to long nights, the rate of degradation is<br />

slow so that the amount accumulated during the short days is<br />

exhausted only at the end of the night. Conversely, in short<br />

© CSIRO 2007 10.1071/FP06313 1445-4408/07/060465


466 Functional Plant Biology S. C. Zeeman et al.<br />

nights, the rate of degradation is high, but again the reserves<br />

are exhausted only at the end of the night (Fig. 1A; Gibon et al.<br />

2004; Lu et al. 2005).<br />

When grown in a diurnal cycle, mutants of Arabidopsis that<br />

cannot synthesise starch, or mutants that can synthesise starch<br />

but cannot remobilise it, grow more slowly than wild type plants<br />

(Caspar et al. 1985, 1991; Zeeman et al. 1998). Caspar et al.<br />

(1985) showed that the reduction in growth rate of a starchless<br />

line (pgm – lacking plastidial phosphoglucomutase) was more<br />

severe in short days than in long days, whereas in continuous<br />

A<br />

Starch content<br />

B<br />

CO2<br />

8-h photoperiod<br />

Short days: a high %<br />

of assimilated C<br />

is partitioned<br />

into starch<br />

Chloroplast stroma<br />

Cytosol<br />

Long days: a low<br />

% of assimilated C is<br />

partitioned into starch<br />

Calvin<br />

cycle<br />

Triosephosphates<br />

Sucrose<br />

light, its growth was equal to that of the wild type. This illustrates<br />

the important role that starch plays in sustaining metabolism<br />

at night.<br />

The pathway of starch synthesis in leaves (Fig. 1B) begins<br />

with the production of the glucosyl donor, ADPglucose<br />

(ADPGlc), from the Calvin cycle intermediate, fructose-6phosphate.<br />

Subsequent production of amylopectin, the major<br />

component of starch, requires the actions of three classes<br />

of enzymes. Starch synthases elongate existing chains by<br />

transferring the glucosyl moiety of ADPGlc to an existing<br />

16-h photoperiod<br />

Short nights:<br />

starch breakdown<br />

is fast<br />

Long nights:<br />

starch breakdown<br />

is slow<br />

0 4 8 12<br />

Time (h)<br />

16 20 24<br />

PGI<br />

PGM AGPase<br />

SS<br />

BE<br />

DBE<br />

Fru6P Glc6P Glc1P ADPGlc<br />

C<br />

ATP P~P i<br />

s<br />

2 P i<br />

Starch<br />

Fig. 1. Starch biosynthesis in Arabidopsis leaves. (A) A scheme illustrating the pattern of<br />

starch accumulation <strong>and</strong> utilisation in Arabidopsis leaves during different diurnal regimes.<br />

The black line represents the near-symmetrical pattern observed under 12-h light/12-h dark<br />

conditions. The rates of starch synthesis <strong>and</strong> starch degradation respond in complementary ways<br />

to short-day (red) or long-day (blue) conditions, as indicated (see also Gibon et al. 2004; Lu et al.<br />

2005). (B) The pathway of starch biosynthesis in the chloroplast stroma. PGI, phosphoglucose<br />

isomerase; PGM, phosphoglucomutase; AGPase, ADPglucose pyrophosphorylase; SS, starch<br />

synthase; BE, branching enzyme; DBE, debranching enzyme; Glc-1-P, glucose-1-phosphate.<br />

(C) Transmission electron micrograph of an Arabidopsis palisade mesophyll cell, showing a<br />

chloroplast containing five starch granules (S). Bar = 1 µm.


Pathways of starch breakdown Functional Plant Biology 467<br />

glucan chain, creating a new α-1,4-bond. Branching enzymes<br />

introduce branch points by cutting an existing α-1,4-linked<br />

chain <strong>and</strong> transferring the cut portion of the chain (6 or<br />

more glucosyl units in length) to an adjacent chain creating<br />

an α-1,6-bond. Debranching enzymes are also required<br />

for amylopectin synthesis, probably by selectively removing<br />

incorrectly positioned branches. The combined actions of<br />

starch synthases, branching enzymes <strong>and</strong> debranching enzymes<br />

produce a glucan in which branch points are non-r<strong>and</strong>omly<br />

distributed. Linear segments of neighbouring chains form double<br />

helices that align into crystalline layers, between which there are<br />

non-crystalline layers containing the branch points (Fig. 1B; for<br />

further details, see Zeeman et al. 2007 <strong>and</strong> references therein).<br />

The resultant starch granule that forms between the thylakoid<br />

membranes in Arabidopsis chloroplasts is an insoluble discoid<br />

structure, 1–2 µm in diameter <strong>and</strong> 0.2–0.5 µm thick (Fig. 1C).<br />

Although the individual functions of the starch biosynthetic<br />

enzymes are reasonably well described, the way in which their<br />

activities are coordinated to generate the specific branching<br />

pattern of amylopectin <strong>and</strong> the mechanism of starch granule<br />

initiation are not well understood.<br />

Rethinking the initial steps in starch breakdown<br />

Mobilisation of the starch granule involves the conversion of<br />

the insoluble, semi-crystalline matrix formed by amylopectin<br />

to soluble sugars that can feed into intermediary metabolism.<br />

During starch breakdown in the endosperm of germinated cereal<br />

seeds, α-amylase (an endoamylase) initiates starch degradation<br />

by releasing from the granule a mixture of soluble glucans<br />

that serve as substrates for other hydrolytic enzymes. The<br />

aleurone layer <strong>and</strong> the scutellum – the living cells surrounding<br />

the endosperm – secrete α-amylase into the non-living, starchcontaining<br />

endosperm. The regulation of α-amylase gene<br />

expression has been extensively studied in the aleurone of barley<br />

(Fincher 1989; Gubler et al. 1995, 1999), <strong>and</strong> repression of<br />

α-amylase gene expression in rice using RNAi has <strong>recent</strong>ly been<br />

shown to delay germination (Asatsuma et al. 2005).<br />

Investigation of α-amylase function in Arabidopsis indicates<br />

that this enzyme has only a minor role in starch metabolism.<br />

The Arabidopsis genome encodes three putative α-amylases<br />

(AMY1–AMY3; Stanley et al. 2002). Arabidopsis seeds store<br />

oil rather than starch, so there is no role for α-amylase in<br />

seedling establishment analogous to that in cereals. However, it<br />

was reasonable to suppose that α-amylase might be involved in<br />

leaf starch degradation. Indeed, consistent with this supposition,<br />

initial biochemical studies of a high starch mutant, sex4<br />

(for starch-excess), identified a deficiency in a chloroplasttargeted<br />

isoform of α-amylase (Zeeman et al. 1998). However,<br />

although subsequent studies have confirmed the deficiency in<br />

the chloroplast α-amylase (AMY3) in sex4, they have also<br />

shown that the SEX4 gene does not encode α-amylase (Yu<br />

et al. 2005; Niittylä et al. 2006) <strong>and</strong> that α-amylase is not<br />

required for starch degradation. Yu et al. (2005) reported that<br />

mutants lacking AMY3 have normal rates of starch degradation.<br />

This was also the case for triple mutants in which all three<br />

α-amylase genes were disrupted. Thus, there is another<br />

mechanism for the initiation of starch degradation in leaves.<br />

Nevertheless, α-amylase isoforms homologous to AMY3 are<br />

widely conserved in the plant kingdom <strong>and</strong> in some situations<br />

endoamylolysis might contribute to starch mobilisation (see<br />

below <strong>and</strong> Delatte et al. 2006).<br />

Glucan phosphorylation –akeystep in starch<br />

breakdown<br />

There is a strong body of evidence indicating that the process of<br />

starch breakdown inside plastids requires the action of glucan,<br />

water dikinase (GWD). This enzyme catalyses the transfer of<br />

the β-phosphate of ATP to the C-6 position of glucosyl residues<br />

of amylopectin (Ritte et al. 2000, 2006). Although the presence<br />

of covalently linked phosphate in some starches has long been<br />

known, its biological significance has come to light only <strong>recent</strong>ly.<br />

The extent of phosphorylation varies considerably, depending on<br />

the source of the starch. In Arabidopsis leaf starch, ∼1 in 2000<br />

glucosyl residues is phosphorylated, whereas potato tuber starch<br />

contains four times as much. Cereal endosperm starch contains<br />

little or no phosphate.<br />

Glucan, water dikinase was discovered by Lorberth et al.<br />

(1998) as a protein binding to starch granules extracted from<br />

potato tubers (the protein was initially named ‘R1’). Antisense<br />

repression of GWD resulted in a reduction in starch-bound<br />

phosphate. It also caused starch to accumulate in leaves<br />

<strong>and</strong> prevented the cold-induced sweetening of stored tubers<br />

(a process during which starch is broken down). Thus, Lorberth<br />

et al. (1998) proposed that the phosphorylation of starch<br />

is necessary for its subsequent mobilisation. Although the<br />

phosphorylation of amylopectin occurs as starch is synthesised<br />

(Nielsen et al. 1994), GWD is reported to bind to the surface<br />

of potato leaf starch granules preferentially at night (Ritte et al.<br />

2000). Furthermore, Ritte et al. (2004) provided evidence that<br />

the rate of phosphorylation of the granule surface is higher<br />

during starch breakdown than during biosynthesis. However,<br />

Mikkelsen et al. (2005) reported that GWD is redox-regulated<br />

<strong>and</strong> that the fraction bound to starch granules is in an inactive,<br />

oxidised form, whereas the soluble form is reduced <strong>and</strong> active.<br />

This observation appears inconsistent with a higher rate of nighttime<br />

phosphorylation.<br />

Yu et al. (2001) reported that the sex1 mutant of Arabidopsis<br />

also lacks GWD (sex1 was initially named TC26, then TC265<br />

by Caspar et al. 1991). Trethewey <strong>and</strong> ap Rees (1994a, 1994b)<br />

conducted biochemical studies of sex1 <strong>and</strong> proposed that the<br />

cause of the sex phenotype was a deficiency in the capacity<br />

of the mutant to export glucose from the chloroplast at night.<br />

However, as the sex phenotype was shown to be due to<br />

the mutation in the gene encoding GWD, this conclusion<br />

seems incorrect. The importance of glucose export for starch<br />

degradation still awaits evaluation. Like the potato plants in<br />

which GWD was repressed, the sex1 mutant has reduced levels<br />

of starch-bound phosphate. In sex1 null mutants, the starch levels<br />

are from 4- to 6-fold higher than in wild type plants at the end<br />

of the day, <strong>and</strong> phosphate levels in the starch are reduced to<br />

the limit of detection (Yu et al. 2001; Ritte et al. 2006). Starch<br />

accumulates gradually in the mutant leaves, with young, newly<br />

emerged leaves containing less than mature leaves (Zeeman <strong>and</strong><br />

ap Rees 1999).<br />

A second protein, which shares sequence similarity with<br />

GWD, is now known to phosphorylate amylopectin exclusively


468 Functional Plant Biology S. C. Zeeman et al.<br />

on the C3 position of the glucosyl residues (Baunsgaard et al.<br />

2005; Kötting et al. 2005; Ritte et al. 2006). This second enzyme<br />

requires that its glucan substrate is already phosphorylated<br />

(hence the name phosphoglucan, water dikinase – PWD) <strong>and</strong><br />

so is dependent on the previous action of GWD. Mutants<br />

of Arabidopsis lacking PWD also exhibit a starch-excess<br />

phenotype, although much less severe than that of mutants<br />

lacking GWD. It is important to note that the mechanism<br />

by which the presence of phosphate groups on amylopectin<br />

permits degradation is still not clear. A plausible hypothesis is<br />

that phosphorylation disrupts the semi-crystalline packing of<br />

the starch granule providing ‘access points’ at which glucan<br />

metabolising enzymes can act on the exposed amylopectin<br />

chains. A high rate of transient phosphorylation at the surface<br />

of the starch granule at night could thus facilitate degradation<br />

(Fig. 2A, B). As PWD acts on glucans that have already been<br />

phosphorylated by GWD, it is reasonable to speculate that their<br />

combined actions result in two or more phosphorylated glucosyl<br />

residues in close proximity to each other. Phosphate pairs or<br />

clusters might be more effective in causing a localised disruption<br />

of amylopectin packing than single phosphate groups. GWD<br />

alone might distribute phosphate groups evenly, resulting in only<br />

partial disruption of the granule surface. This could restrict the<br />

rate of starch degradation <strong>and</strong> explain the mild starch-excess<br />

phenotype of the pwd mutant. However, there is as yet no direct<br />

evidence to support such a model.<br />

It is not currently known how or when the phosphate groups<br />

added by GWD <strong>and</strong> PWD are removed from the glucans during<br />

starch breakdown. Phosphoglucans may be released from the<br />

granule <strong>and</strong> metabolised in the chloroplast stroma. Alternatively,<br />

phosphate groups may be removed directly from the granule<br />

surface. One possibility is that the <strong>recent</strong>ly discovered glucanbinding<br />

phosphatase encoded at the SEX4 locus fulfils this<br />

role. Like sex1, sex4 was identified through screening for<br />

mutant plants that contain starch at the end of the night,<br />

when the wild type has exhausted its starch reserves. The<br />

mutant has a reduced rate of starch degradation at night <strong>and</strong><br />

accumulates starch as a result of the imbalance between daytime<br />

synthesis <strong>and</strong> night-time breakdown. Mature sex4 leaves<br />

have three times as much starch as the wild type at the end<br />

of the day (Zeeman et al. 1998; Zeeman <strong>and</strong> ap Rees 1999).<br />

Niittylä et al. (2006) showed that SEX4 encodes a predicted dualspecificity<br />

protein phosphatase (DSP), initially described by<br />

Fordham-Skelton et al. (2002) as PTPKIS (for protein tyrosine<br />

phosphatase with a kinase interaction sequence). More <strong>recent</strong>ly<br />

the same protein has been described by Kerk et al. (2006) as<br />

DSP <strong>and</strong> Sokolov et al. (2006) as DSP4. The domain of SEX4<br />

initially described as a KIS domain is now established as a<br />

carbohydrate-binding module (Kerk et al. 2006; Niittylä et al.<br />

2006; Sokolov et al. 2006). Remarkably, SEX4 has homologues<br />

in animals (called laforin), mutation of which results in Lafora<br />

disease, a neurodegenerative condition caused by abnormal<br />

glycogen metabolism. Worby et al. (2006) reported that laforin<br />

can mediate the dephosphorylation of a branched glucan<br />

(potato amylopectin) in vitro <strong>and</strong> stated that SEX4 had similar<br />

capabilities. If further studies confirm that phosphoglucans<br />

are indeed the substrate for SEX4, it will be necessary to<br />

determine whether the enzymes substrate in vivo is the granule<br />

Granule surface Stroma<br />

BAM<br />

GWD<br />

PWD<br />

BAM<br />

ISA3<br />

ISA3<br />

ATP<br />

AMP + P i<br />

ATP<br />

AMP + P i<br />

A<br />

B<br />

C<br />

Maltose<br />

D<br />

Maltooligosaccharides<br />

Fig. 2. A hypothetical mechanism for the initiation of starch granule<br />

breakdown involving amylopectin phosphorylation. (A) The granule surface<br />

is depicted as a semi-crystalline layer of amylopectin. Grey boxes represent<br />

double helices formed by adjacent unbranched glucan chains (inset).<br />

(B) Glucan, water dikinase (GWD) phosphorylates the surface of the<br />

starch granule (red circle). Phosphoglucan, water dikinase (PWD) adds a<br />

second phosphate group (gold circle), probably very close to one added<br />

by GWD. Phosphorylation might disrupt the packing of the double helices<br />

of amylopectin. (C) β-Amylase (BAM) degrades accessible, free chains at<br />

the granule surface, liberating maltose, but cannot work past branch points.<br />

(D) The debranching enzyme isoamylase 3 (ISA3) removes short branches,<br />

liberating short malto-oligosaccharides <strong>and</strong> enabling the continued action of<br />

β-amylase.


Pathways of starch breakdown Functional Plant Biology 469<br />

surface or a pool of soluble phosphoglucans. Furthermore,<br />

testable models are required to explain why phosphate groups<br />

first need to be added <strong>and</strong> subsequently removed for normal<br />

starch breakdown to occur. It remains possible that SEX4<br />

in fact dephosphorylates <strong>and</strong> thus regulates the activities of<br />

proteins involved in starch degradation as well as, or instead<br />

of, dephosphorylating amylopectin.<br />

The hydrolysis of glucans at the starch granule surface<br />

Given that α-amylase is not required for starch breakdown<br />

in Arabidopsis, other enzymes must be able to attack the<br />

starch granule. Results from several laboratories indicate<br />

that β-amylase (an exoamylase which produces maltose) <strong>and</strong><br />

debranching enzymes may fulfil this role by acting in synergy.<br />

Scheidig et al. (2002) first reported that antisense repression of a<br />

chloroplast-targeted isoform of β-amylase (CT-BMY1) in potato<br />

caused an accumulation of starch in leaves. Kaplan <strong>and</strong> Guy<br />

(2005) showed that the repression of the Arabidopsis homologue<br />

of CT-BMY (BAM3, also called BMY8) using RNAi also caused<br />

a sex phenotype, a result supported by our analysis of a bam3<br />

null mutant (S. C. Zeeman, A. M. Smith <strong>and</strong> S. M. Smith,<br />

unpubl. data). Scheidig et al. (2002) analysed the activity of the<br />

potato protein in vitro <strong>and</strong> showed that, in addition to activity<br />

on soluble substrates, PCT-BMY1 was able to liberate a small<br />

amount of maltose from isolated potato starch granules. Such<br />

experiments can be difficult to interpret. On the one h<strong>and</strong>, a<br />

low activity of β-amylase against intact starch granules is not<br />

surprising as only the external linear chains of the granule will<br />

be available for hydrolysis. β-amylase cannot work past the<br />

branch points of amylopectin <strong>and</strong> would rapidly create a β-limit<br />

structure consisting of short branches or ‘stubs’ on the granule<br />

surface. The action of a debranching enzyme would be necessary<br />

for further β-amylolysis to occur. Furthermore, if β-amylase<br />

requires that amylopectin is first phosphorylated by GWD,<br />

in vitro conditions may be unsuitable to show its true activity.<br />

On the other h<strong>and</strong>, it can be difficult to purify starch granules<br />

that are free of contaminating proteins (e.g. α-amylases), which<br />

might provide soluble substrates for β-amylase.<br />

By analysing the lengths of the glucan chains on the<br />

outermost layer of potato starch granules, Ritte et al. (2004)<br />

showed that the abundance of short chains 3–5 glucose residues<br />

in length increases markedly in the first hour of the night. This<br />

observation is consistent with β-amylolytic attack of the external<br />

chains. The short branches that result are probably removed<br />

directly from the granule surface by the debranching enzyme<br />

isoamylase 3 (ISA3; Fig. 2C, D). Loss of this enzyme causes<br />

starch to accumulate in Arabidopsis leaves, implicating it in<br />

starch breakdown (Wattebled et al. 2005; Delatte et al. 2006).<br />

Analysis of the activity of the potato ISA3 protein in vitro showed<br />

that it was particularly active on β-limit dextrins, <strong>and</strong> also had<br />

some action on isolated starch granules (Hussain et al. 2003).<br />

Delatte et al. (2006) analysed the isa3 mutant phenotype of<br />

Arabidopsis in detail <strong>and</strong> showed that the plants have a reduced<br />

rate of starch degradation compared to the wild type <strong>and</strong> that<br />

short chains 3–5 glucose residues in length are abundant in the<br />

starch. This <strong>suggest</strong>s that short chains are not efficiently removed<br />

by debranching in the isa3 mutant. Furthermore, expression of<br />

an ISA3-GFP construct in Arabidopsis leaf protoplasts showed<br />

that the fluorescent fusion protein localises to the starch granule<br />

surface, consistent with its proposed role there (Delatte et al.<br />

2006). Taken together, these data indicate that the surface of<br />

the starch granule is subject to progressive exoamylolysis <strong>and</strong><br />

debranching <strong>and</strong> that this process may be dependent on the<br />

previous phosphorylation of the granule to disrupt the packing<br />

of double helices (Fig. 2).<br />

The mechanism described above results in the release<br />

of maltose together with some linear malto-oligosaccharides<br />

directly from the granule surface. Interestingly, in mutants<br />

lacking components of this pathway (e.g. isa3 <strong>and</strong> bam3),<br />

starch degradation still occurs, albeit at a reduced rate. This<br />

shows that other enzymes are also able to degrade starch.<br />

Delatte et al. (2006) investigated the effect of mutating a<br />

second debranching enzyme, limit dextrinase (LDA), alone<br />

<strong>and</strong> in addition to ISA3. The lda mutant displays normal<br />

starch metabolism indicating that, like α-amylase, LDA is not<br />

required for normal rates of starch breakdown. Interestingly,<br />

when isa3 <strong>and</strong> lda mutations were combined, the double mutants<br />

exhibited a more severe starch-excess phenotype than the isa3<br />

single mutant, showing that LDA was contributing to starch<br />

breakdown in the isa3 background. Further analysis of isa3<br />

<strong>and</strong> of the isa3/lda double mutant provided evidence that<br />

α-amylase might also contribute to starch degradation in these<br />

lines. First, there was still appreciable starch degradation in<br />

the double mutant. Second, the amount of the chloroplastic<br />

α-amylase AMY3 was increased in both lines relative to the<br />

wild type, as was total endoamylolytic activity. Third, branched<br />

oligosaccharides accumulated in the double mutant, but not in<br />

the isa3 single mutant. Delatte et al. (2006) interpreted these<br />

data to mean that in isa3, starch degradation proceeds at least<br />

partly through α-amylase. Liberated branched oligosaccharides<br />

are then debranched by LDA in the stroma to yield linear<br />

oligosaccharides, which can be metabolised further. In the<br />

isa3/lda double mutant, branched oligosaccharides that are<br />

released by AMY3 cannot be metabolised <strong>and</strong> so accumulate<br />

in the stroma. Thus, there may be two <strong>distinct</strong> <strong>mechanisms</strong> by<br />

which the surface of the starch granule is degraded: a major<br />

pathway involving β-amylase <strong>and</strong> ISA3, <strong>and</strong> a minor pathway<br />

involving α-amylase <strong>and</strong> LDA (Fig. 3).<br />

Metabolism of malto-oligosaccharides<br />

There is good evidence that the major product of starch<br />

breakdown is maltose, which is exported by a unique chloroplast<br />

envelope transporter, MEX1, to the cytosol (Niittylä et al. 2004;<br />

Weise et al. 2004). In the cytosol, maltose is metabolised by<br />

a glucosyltransferase, DPE2 (Chia et al. 2004; Lu <strong>and</strong> Sharkey<br />

2004). In the reaction catalysed by DPE2, one glucosyl residue is<br />

transferred to an acceptor molecule, whereas the other is released<br />

as glucose. Glycogen, amylopectin <strong>and</strong> oligosaccharides will<br />

all serve as acceptors for the transferred glucosyl residue in<br />

vitro (Chia et al. 2004; Lloyd et al. 2004; Fettke et al. 2006).<br />

However, the likely acceptor in vivo is a soluble heteroglycan<br />

present in the cytosol (Fettke et al. 2005a). Glucose liberated<br />

by DPE2 can be phosphorylated by hexokinase <strong>and</strong> thereby<br />

enter the hexose-phosphate pool, whereas the glucosyl residue


470 Functional Plant Biology S. C. Zeeman et al.<br />

Chloroplast<br />

metabolism<br />

Glc1P<br />

AMY3<br />

Branched maltooligosaccharides<br />

LDA<br />

Starch<br />

ISA3<br />

Linear maltooligosaccharides<br />

PHS1<br />

β-amylase<br />

pGlcT<br />

DPE1<br />

Glucose<br />

Maltose<br />

β-amylase<br />

(BAM3)<br />

MEX1<br />

Chloroplast stroma<br />

Cytosol<br />

Fig. 3. The <strong>pathways</strong> of starch breakdown in Arabidopsis leaves. Maltose is the major product of starch<br />

breakdown, exported to the cytosol by the chloroplast envelope transporter, MEX1. Maltose is produced by<br />

β-amylase acting either at the granule surface (see Fig. 2) or on soluble malto-oligosaccharides. The actions of<br />

isoamylase (ISA3, see Fig. 2) or α-amylase (AMY3) on the granule may release soluble malto-oligosaccharides,<br />

which, if branched, can be debranched by limit dextrinase (LDA). Short linear oligosaccharides (e.g. maltotriose)<br />

can be metabolised by disproportionating enzyme (DPE1) liberating glucose for export to the cytosol (by the<br />

glucose transporter, pGlcT) <strong>and</strong> larger malto-oligosaccharides for continued degradation. Chloroplast α-glucan<br />

phosphorylase (PHS1) can use linear oligosaccharides to produce Glc-1-P for chloroplast metabolism. Some steps<br />

are depicted with dashed lines because removal of these enzymes individually does not prevent a normal rate of<br />

starch breakdown at night.<br />

transferred to the cytosolic acceptor might be liberated as<br />

glucose-1-phosphate by the cytosolic α-glucan phosphorylase,<br />

PHS2 (Lu et al. 2006a).<br />

The loss of MEX1 or DPE2 leads to the massive accumulation<br />

of maltose (Chia et al. 2004; Lu <strong>and</strong> Sharkey 2004; Niittylä et al.<br />

2004). Lu et al. (2006b) confirmed that in the mex1 mutant,<br />

the maltose accumulates in the chloroplast, whereas in the dpe2<br />

mutant it accumulates in both the chloroplast <strong>and</strong> the cytosol.<br />

These data are consistent with the idea that MEX1 facilitates<br />

bi-directional diffusion of maltose across the envelope, <strong>and</strong><br />

that DPE2 is cytosolic. These phenotypes also <strong>suggest</strong> that<br />

α-glucosidase (also known as maltase) does not have a<br />

significant role in maltose metabolism in Arabidopsis leaves.<br />

In potato, there remains some uncertainty over the cellular<br />

location of DPE2. Initially, Lloyd et al. (2004) reported the<br />

protein to be localised in the chloroplast. This implies a different<br />

mechanism of maltose metabolism. However, a subsequent study<br />

has indicated that potato DPE2 is cytosolic, as in Arabidopsis<br />

(Fettke et al. 2005b). For additional reviews on cytosolic maltose<br />

metabolism, see Smith et al. (2005), Lu <strong>and</strong> Sharkey (2006) <strong>and</strong><br />

Fettke et al. (2007).<br />

In addition to direct release from the granule surface, maltose<br />

can be produced by the β-amylolysis of linear oligosaccharides<br />

in the chloroplast stroma. Linear oligosaccharides themselves<br />

can be released from the granule by debranching enzyme<br />

(i.e. ISA3), by α-amylase, or by the debranching of soluble<br />

branched glucans released by α-amylase. Early biochemical<br />

studies showed that β-amylase is able to efficiently metabolise<br />

oligosaccharides longer than 4 glucosyl residues in length,<br />

but that maltotriose is a poor substrate (Chapman et al.<br />

1972). Thus, in addition to maltose, some maltotriose will be<br />

produced when β-amylase metabolises linear chains with an<br />

odd number of glucosyl residues. Furthermore, ISA3 might<br />

release maltotriosyl branches from the starch granule surface<br />

(Delatte et al. 2006).<br />

Maltotriose is metabolised in the chloroplast stroma by<br />

disproportionating enzyme (D-enzyme, DPE1). The loss of<br />

D-enzyme in Arabidopsis results in maltotriose accumulation<br />

during starch breakdown (Critchley et al. 2001). When provided<br />

solely with maltotriose in vitro, D-enzyme transfers a maltosyl<br />

moiety from one molecule to another, producing maltopentaose<br />

<strong>and</strong> glucose (Jones <strong>and</strong> Whelan 1969). In vivo, glucose produced<br />

this way could be exported from the chloroplast by the<br />

glucose transporter (Schäfer et al. 1977; Weber et al. 2000),<br />

<strong>and</strong> maltopentaose could serve as a substrate for β-amylase,<br />

resulting in maltose <strong>and</strong> maltotriose. Such a cycle would result


Pathways of starch breakdown Functional Plant Biology 471<br />

in the net conversion of maltotriose to equimolar amounts of<br />

maltose <strong>and</strong> glucose. However, D-enzyme can catalyse a wide<br />

range of reactions <strong>and</strong> it is likely that it acts on other maltooligosaccharides,<br />

as well as maltotriose in vivo. Thus, it might<br />

function to recycle short malto-oligosaccharides in general.<br />

Interestingly, D-enzyme does not produce or metabolise maltose<br />

(Jones <strong>and</strong> Whelan 1969; Lin <strong>and</strong> Preiss 1988). This means that<br />

when maltose is produced during starch breakdown, it is not<br />

re-incorporated into the malto-oligosaccharide pool <strong>and</strong> is only<br />

metabolised after export to the cytosol.<br />

In addition to the production of the free sugars maltose<br />

<strong>and</strong> glucose, glucose-1-phosphate can be produced from<br />

linear malto-oligosaccharides by the chloroplastic form of<br />

α-glucan phosphorylase (Fig. 3). The role of phosphorolysis<br />

in the chloroplast is not yet clear, but it probably represents<br />

a pathway for the production of substrates for chloroplast<br />

metabolism, as the chloroplast envelope is not permeable to<br />

hexose-phosphates. Repression of chloroplastic phosphorylase<br />

by antisense techniques in potato had no detectable effect on<br />

plant growth or starch metabolism (Sonnewald et al. 1995).<br />

Complete removal of the enzyme (PHS1) by insertional<br />

mutagenesis in Arabidopsis had only a minor effect on leaf<br />

starch metabolism (Zeeman et al. 2004). In phs1 mutants,<br />

small groups of cells bordering necrotic lesions were found to<br />

contain large amounts of starch. Necrotic lesions were more<br />

frequently observed in phs1 mutants than in wild type plants,<br />

<strong>and</strong> phs1 plants were particularly susceptible to acute water<br />

<strong>and</strong> salt stresses. Plausible roles for phosphorylase include the<br />

provision of substrates for the chloroplastic oxidative pentose<br />

phosphate pathway (Zeeman et al. 2004). This pathway might<br />

be important during the dark for the provision of reducing<br />

agents necessary to tolerate stress conditions (e.g. for the<br />

scavenging of reactive oxygen species). Phosphorylase may also<br />

supply carbon that can replenish intermediates of the Calvin<br />

cycle. This may be particularly important under photorespiratory<br />

conditions where such intermediates could become depleted<br />

(Weise et al. 2006).<br />

Conclusions<br />

The findings summarised in this mini-review illustrate<br />

both the progress made <strong>recent</strong>ly in underst<strong>and</strong>ing starch<br />

degradation, <strong>and</strong> the number of major questions that remain<br />

unanswered. Although the evidence supports the existence<br />

of <strong>distinct</strong> <strong>mechanisms</strong> of starch granule degradation in<br />

Arabidopsis – a minor endoamylolytic pathway <strong>and</strong> a<br />

major exoamylolytic/debranching pathway – the functional<br />

<strong>distinct</strong>ions between the two <strong>pathways</strong> is not clear. The degrees<br />

to which the results from Arabidopsis can be extended to<br />

other species also remain unclear. As starch is the form<br />

in which carbohydrate is stored in leaves, it seems likely<br />

that its metabolism may respond to diverse physiological <strong>and</strong><br />

environmental parameters. The narrow range of environmental<br />

conditions typically used in laboratory experiments may not<br />

bring all such parameters to the fore. It is particularly important<br />

that further progress is made in underst<strong>and</strong>ing how the<br />

phosphorylation of amylopectin influences starch breakdown.<br />

The severe phenotype that results from the loss of GWD implies<br />

that it exerts significant control over the process. It will be<br />

very interesting to determine how the phosphatase SEX4 is<br />

involved <strong>and</strong> to discover the extent of the similarities between<br />

starch metabolism <strong>and</strong> animal glycogen metabolism. Finally, a<br />

complete underst<strong>and</strong>ing of the <strong>pathways</strong> of starch breakdown<br />

will enable us to study how they are regulated <strong>and</strong> integrated<br />

with the rest of plant primary metabolism.<br />

Acknowledgements<br />

We offer particular thanks to our close collaborators, Professor Alison<br />

M. Smith <strong>and</strong> Professor Steven M. Smith, <strong>and</strong> to our colleagues in the<br />

Arabidopsis Starch Metabolism Network (www.starchmetnet.org, accessed<br />

17 March 2007). We also thank Martine Trevisan <strong>and</strong> Simona Eicke for<br />

excellent technical support, past <strong>and</strong> present, in our laboratory. Our research<br />

is funded by the Roche Research Foundation, the Swiss National Science<br />

Foundation (Nation Centre for Competence in Research – Plant Survival <strong>and</strong><br />

Grant 3100–067312.01/1), <strong>and</strong> the EMBO Young Investigator Programme.<br />

References<br />

Asatsuma S, Sawada C, Itoh K, Okito M, Kitajima A, Mitsui T (2005)<br />

Involvement of α-amylase I-1 in starch degradation in rice chloroplasts.<br />

Plant & Cell Physiology 46, 858–869. doi: 10.1093/pcp/pci091<br />

Baunsgaard L, Lütken H, Mikkelsen R, Glaring MA, Pham TT, Blennow A<br />

(2005) A <strong>novel</strong> isoform of glucan, water dikinase phosphorylates<br />

pre-phosphorylated α-glucans <strong>and</strong> is involved in starch degradation<br />

in Arabidopsis. The Plant Journal 41, 595–605. doi: 10.1111/j.1365-<br />

313X.2004.02322.x<br />

Buléon A, Colonna P, Planchot V, Ball S (1998) Starch granules: structure<br />

<strong>and</strong> biosynthesis. International Journal of Biological Macromolecules<br />

23, 85–112. doi: 10.1016/S0141-8130(98)00040-3<br />

Caspar T, Huber SC, Somerville C (1985) Alterations in growth,<br />

photosynthesis, <strong>and</strong> respiration in a starchless mutant of Arabidopsis<br />

thaliana (L.) deficient in chloroplast phosphoglucomutase activity.<br />

Plant Physiology 79, 11–17.<br />

Caspar T, Lin T-P, Kakefuda G, Benbow L, Preiss J, Somerville C (1991)<br />

Mutants of Arabidopsis with altered regulation of starch degradation.<br />

Plant Physiology 95, 1181–1188.<br />

Chapman GW Jr, Pallas JE Jr, Mendicino J (1972) The hydrolysis of<br />

maltodextrins by a β-amylase isolated from leaves of Vicia faba.<br />

Biochimica et Biophysica Acta 276, 491–507. doi: 10.1016/0005-<br />

2744(72)91010-8<br />

Chatterton NJ, Silvius JE (1980) Photosynthate partitioning into leaf<br />

starch as affected by daily photosynthetic period duration in 6<br />

species. Physiologia Plantarum 49, 141–144. doi: 10.1111/j.1399-<br />

3054.1980.tb02642.x<br />

Chia T, Thorneycroft D, Chapple A, Messerli G, Chen J, Zeeman SC,<br />

Smith SM, Smith AM (2004) A cytosolic glucosyltransferase is<br />

required for conversion of starch to sucrose in Arabidopsis leaves<br />

at night. The Plant Journal 37, 853–863. doi: 10.1111/j.1365-<br />

313X.2003.02012.x<br />

Critchley JH, Zeeman SC, Takaha T, Smith AM, Smith SM (2001) A critical<br />

role for disproportionating enzyme in starch breakdown is revealed by<br />

a knock-out mutation in Arabidopsis. The Plant Journal 26, 89–100.<br />

doi: 10.1046/j.1365-313x.2001.01012.x<br />

Delatte T, Umhang M, Trevisan M, Eicke S, Thorneycroft D, Smith SM,<br />

Zeeman SC (2006) Evidence for <strong>distinct</strong> <strong>mechanisms</strong> of starch granule<br />

breakdown in plants. The Journal of Biological Chemistry 281,<br />

12 050–12 059. doi: 10.1074/jbc.M513661200<br />

Fettke J, Eckermann N, Tiessen A, Geigenberger P, Steup M (2005a)<br />

Identification, subcellular localization <strong>and</strong> biochemical characterization<br />

of water-soluble heteroglycans (SHG) in leaves of Arabidopsis<br />

thaliana L.: <strong>distinct</strong> SHG reside in the cytosol <strong>and</strong> in the<br />

apoplast. The Plant Journal 43, 568–585. doi: 10.1111/j.1365-<br />

313X.2005.02475.x


472 Functional Plant Biology S. C. Zeeman et al.<br />

Fettke J, Poeste S, Eckermann N, Tiessen A, Pauly M, Geigenberger P,<br />

Steup M (2005b) Analysis of cytosolic heteroglycans from leaves<br />

of transgenic potato (Solanum tuberosum L.) plants that under- or<br />

overexpress the Pho 2 phosphorylase isozyme. Plant & Cell Physiology<br />

46, 1987–2004. doi: 10.1093/pcp/pci214<br />

Fettke J, Chia T, Eckermann N, Smith A, Steup M (2006) A transglucosidase<br />

necessary for starch degradation <strong>and</strong> maltose metabolism in leaves at<br />

night acts on cytosolic heteroglycans (SHG). The Plant Journal 46,<br />

668–684. doi: 10.1111/j.1365-313X.2006.02732.x<br />

Fettke J, Eckermann N, Kötting O, Ritte G, Steup M (2007) Novel starchrelated<br />

enzymes <strong>and</strong> carbohydrates. Cellular <strong>and</strong> Molecular Biology 52,<br />

OL883–OL904.<br />

Fincher GB (1989) Molecular <strong>and</strong> cellular biology associated with<br />

endosperm mobilization in germinating cereal grains. Annual Review<br />

of Plant Physiology <strong>and</strong> Plant Molecular Biology 40, 305–346.<br />

doi: 10.1146/annurev.pp.40.060189.001513<br />

Fordham-Skelton AP, Chilley P, Lumbreras V, Reignoux S, Fenton TR,<br />

Dahm CC, Pages M, Gatehouse JA (2002) A <strong>novel</strong> higher plant protein<br />

tyrosine phosphatase interacts with SNF1-related protein kinases via<br />

a KIS (kinase interaction sequence) domain. The Plant Journal 29,<br />

705–715. doi: 10.1046/j.1365-313X.2002.01250.x<br />

Gibon Y, Bläsing OE, Palacios-Rojas N, Pankovic D, Hendriks JHM,<br />

Fisahn J, Höhne M, Günther M, Stitt M (2004) Adjustment of<br />

diurnal starch turnover to short days: depletion of sugar during the<br />

night leads to a temporary inhibition of carbohydrate utilization,<br />

accumulation of sugars <strong>and</strong> post-translational activation of ADP-glucose<br />

pyrophosphorylase in the following light period. The Plant Journal 39,<br />

847–862. doi: 10.1111/j.1365-313X.2004.02173.x<br />

Gubler F, Kalla R, Roberts JK, Jacobsen JV (1995) Gibberellin-regulated<br />

expression of a MYB gene in barley aleurone cells – evidence for MYB<br />

transactivation of a high-pI α-amylase gene promoter. The Plant Cell 7,<br />

1879–1891. doi: 10.1105/tpc.7.11.1879<br />

Gubler F, Raventos D, Keys M, Watts R, Mundy J, Jacobsen JV (1999) Target<br />

genes <strong>and</strong> regulatory domains of the GAMYB transcriptional activator<br />

in cereal aleurone. The Plant Journal 17, 1–9. doi: 10.1046/j.1365-<br />

313X.1999.00346.x<br />

Hussain H, Mant A, Seale R, Zeeman SC, Hinchliffe E et al. (2003)<br />

Three isoforms of isoamylase contribute different catalytic properties<br />

for the debranching of potato glucans. The Plant Cell 15, 133–149.<br />

doi: 10.1105/tpc.006635<br />

Jones G, Whelan WJ (1969) The action pattern of D-enzyme, a<br />

transmaltodextrinylase from potato. Carbohydrate Research 9, 483–490.<br />

doi: 10.1016/S0008-6215(00)80033-6<br />

Kaplan F, Guy CL (2005) RNA interference of Arabidopsis β-amylase<br />

8 prevents maltose accumulation upon cold shock <strong>and</strong> increases<br />

sensitivity of PSII photochemical efficiency to freezing stress. The Plant<br />

Journal 44, 730–743. doi: 10.1111/j.1365-313X.2005.02565.x<br />

Kerk D, Conley TR, Rodriguez FA, Tran HT, Nimick M, Muench DG,<br />

Moorhead GBG (2006) A chloroplast-localized dual-specificity protein<br />

phosphatase in Arabidopsis contains a phylogenetically dispersed <strong>and</strong><br />

ancient carbohydrate-binding domain, which binds the polysaccharide<br />

starch. The Plant Journal 46, 400–413. doi: 10.1111/j.1365-<br />

313X.2006.02704.x<br />

Kötting O, Pusch K, Tiessen A, Geigenberger P, Steup M, Ritte G<br />

(2005) Identification of a <strong>novel</strong> enzyme required for starch metabolism<br />

in Arabidopsis leaves. The phosphoglucan, water dikinase. Plant<br />

Physiology 137, 242–252. doi: 10.1104/pp.104.055954<br />

Lin T-P, Preiss J (1988) Characterisation of D-enzyme (4-αglucanotransferase)<br />

in Arabidopsis leaf. Plant Physiology 86, 260–265.<br />

Lloyd JR, Blennow A, Burhenne K, Kossmann J (2004) Repression<br />

of a <strong>novel</strong> isoform of disproportionating enzyme (stDPE2) in<br />

potato leads to inhibition of starch degradation in leaves but not<br />

tubers stored at low temperature. Plant Physiology 134, 1347–1354.<br />

doi: 10.1104/pp.103.038026<br />

Lorberth R, Ritte G, Willmitzer L, Kossmann J (1998) Inhibition of a<br />

starch-granule-bound protein leads to modified starch <strong>and</strong> repression<br />

of cold sweetening. Nature Biotechnology 16, 473–477. doi: 10.1038/<br />

nbt0598-473<br />

Lu Y, Sharkey TD (2004) The role of amylomaltase in maltose<br />

metabolism in the cytosol of photosynthetic cells. Planta 218, 466–473.<br />

doi: 10.1007/s00425-003-1127-z<br />

Lu Y, Gehan JP, Sharkey TD (2005) Daylength <strong>and</strong> circadian effects on starch<br />

degradation <strong>and</strong> maltose metabolism. Plant Physiology 138, 2280–2291.<br />

doi: 10.1104/pp.105.061903<br />

Lu Y, Sharkey TD (2006) The importance of maltose in transitory<br />

starch breakdown. Plant, Cell & Environment 29, 353–366.<br />

doi: 10.1111/j.1365-3040.2005.01480.x<br />

Lu Y, Steichen JM, Yao J, Sharkey TD (2006a) The role of cytosolic<br />

α-glucan phosphorylase in maltose metabolism <strong>and</strong> the comparison of<br />

amylomaltase in Arabidopsis <strong>and</strong> Escherichia coli. Plant Physiology<br />

142, 878–889. doi: 10.1104/pp.106.086850<br />

Lu Y, Steichen JM, Weise SE, Sharkey TD (2006b) Cellular <strong>and</strong> organ<br />

level localization of maltose in maltose-excess Arabidopsis mutants.<br />

Planta 224, 935–943. doi: 10.1007/s00425-006-0263-7<br />

Mikkelsen R, Mutenda KE, Mant A, Schürmann P, Blennow A (2005)<br />

α-Glucan, water dikinase (GWD): a plastidic enzyme with redoxregulated<br />

<strong>and</strong> coordinated catalytic activity <strong>and</strong> binding affinity.<br />

Proceedings of the National Academy of Sciences USA 102, 1785–1790.<br />

doi: 10.1073/pnas.0406674102<br />

Nielsen TH, Wischmann B, Enevoldsen K, Møller BL (1994) Starch<br />

phosphorylation in potato tubers proceeds concurrently with de novo<br />

biosynthesis of starch. Plant Physiology 105, 111–117.<br />

Niittylä T, Messerli G, Trevisan M, Chen J, Smith AM, Zeeman SC<br />

(2004) A previously unknown maltose transporter essential for starch<br />

degradation in leaves. Science 303, 87–89. doi: 10.1126/science.<br />

1091811<br />

Niittylä T, Comparot-Moss S, Lue W-L, Messerli G, Trevisan M et al. (2006)<br />

Similar protein phosphatases control starch metabolism in plants <strong>and</strong><br />

glycogen metabolism in mammals. Journal of Biological Chemistry 281,<br />

11 815–11 818. doi: 10.1074/jbc.M600519200<br />

Ritte G, Lorberth R, Steup M (2000) Reversible binding of the starch-related<br />

R1 protein to the surface of transitory starch granules. The Plant Journal<br />

21, 387–391. doi: 10.1046/j.1365-313x.2000.00683.x<br />

Ritte G, Scharf A, Eckermann N, Haebel S, Steup M (2004) Phosphorylation<br />

of transitory starch is increased during degradation. Plant Physiology<br />

135, 2068–2077. doi: 10.1104/pp.104.041301<br />

Ritte G, Heydenreich M, Mahlow S, Haebel S, Kötting O, Steup M<br />

(2006) Phosphorylation of C6- <strong>and</strong> C3-positions of glucosyl residues in<br />

starch is catalysed by <strong>distinct</strong> dikinases. FEBS Letters 580, 4872–4876.<br />

doi: 10.1016/j.febslet.2006.07.085<br />

Schäfer G, Heber U, Heldt HW (1977) Glucose transport into spinach<br />

chloroplasts. Plant Physiology 60, 286–289.<br />

Scheidig A, Fröhlich A, Schulze S, Lloyd JR, Kossmann J (2002)<br />

Down regulation of a chloroplast-targeted β-amylase leads to a<br />

starch-excess phenotype in leaves. The Plant Journal 30, 581–591.<br />

doi: 10.1046/j.1365-313X.2002.01317.x<br />

Smith AM, Zeeman SC, Smith SM (2005) Starch degradation.<br />

Annual Review of Plant Biology 56, 73–98. doi: 10.1146/annurev.<br />

arplant.56.032604.144257<br />

Sokolov LN, Dominguez-Solis JR, Allary A-L, Buchanan BB, Luan S<br />

(2006) A redox-regulated chloroplast protein phosphatase binds to<br />

starch diurnally <strong>and</strong> functions in its accumulation. Proceedings<br />

of the National Academy of Sciences USA 103, 9732–9737.<br />

doi: 10.1073/pnas.0603329103<br />

Sonnewald U, Basner A, Greve B, Steup M (1995) A second L-type<br />

isozyme of potato glucan phosphorylase: cloning, antisense inhibition<br />

<strong>and</strong> expression analysis. Plant Molecular Biology 27, 567–576.<br />

doi: 10.1007/BF00019322


Pathways of starch breakdown Functional Plant Biology 473<br />

Stanley D, Fitzgerald AM, Farnden KJF, McRae EA (2002) Characterization<br />

of putative amylases from apple (Malus domestica) <strong>and</strong> Arabidopsis<br />

thaliana. Biologia 57(Supp), 137–148.<br />

Trethewey RN, ap Rees T (1994a) A mutant of Arabidopsis thaliana<br />

lacking the ability to transport glucose across the chloroplast envelope.<br />

The Biochemical Journal 301, 449–454.<br />

Trethewey RN, ap Rees T (1994b) The role of the hexose transporter<br />

in the chloroplasts of Arabidopsis thaliana L. Planta 195, 168–174.<br />

doi: 10.1007/BF00199675<br />

Wattebled F, Dong Y, Dumez S, Delvallé D, Planchot R et al. (2005)<br />

Mutants of Arabidopsis lacking a chloroplastic isoamylase accumulate<br />

phytoglycogen <strong>and</strong> an abnormal form of amylopectin. Plant Physiology<br />

138, 184–195. doi: 10.1104/pp.105.059295<br />

Weber A, Servaites JC, Geiger DR, Kofler H, Hille D, Gröner F, Hebbeker U,<br />

Flügge U-I (2000) Identification, purification, <strong>and</strong> molecular cloning of<br />

a putative plastidic glucose translocator. The Plant Cell 12, 787–801.<br />

doi: 10.1105/tpc.12.5.787<br />

Weise SE, Weber APM, Sharkey TD (2004) Maltose is the major form of<br />

carbon exported from the chloroplast at night. Planta 218, 474–482.<br />

doi: 10.1007/s00425-003-1128-y<br />

Weise SE, Schrader SM, Kleinbeck KR, Sharkey TD (2006) Carbon<br />

balance <strong>and</strong> circadian regulation of hydrolytic <strong>and</strong> phosphorolytic<br />

breakdown of transitory starch. Plant Physiology 141, 879–886.<br />

doi: 10.1104/pp.106.081174<br />

Worby CA, Gentry MS, Dixon JE (2006) Laforin: a dual specificity<br />

phosphatase that dephosphorylates complex carbohydrates.<br />

Journal of Biological Chemistry 281, 30 412–30 418. doi: 10.1074/<br />

jbc.M606117200<br />

http://www.publish.csiro.au/journals/fpb<br />

Yu T-S, Kofler H, Häusler RE, Hille D, Flügge U-I et al. (2001)<br />

SEX1 is a general regulator of starch degradation in plants <strong>and</strong> not<br />

the chloroplast hexose transporter. The Plant Cell 13, 1907–1918.<br />

doi: 10.1105/tpc.13.8.1907<br />

Yu T-S, Zeeman SC, Thorneycroft D, Fulton DC, Dunstan H et al. (2005)<br />

α-Amylase is not required for breakdown of transitory starch in<br />

Arabidopsis leaves. Journal of Biological Chemistry 280, 9773–9779.<br />

doi: 10.1074/jbc.M413638200<br />

Zeeman SC, Northrop F, Smith AM, ap Rees T (1998) A starch-accumulating<br />

mutant of Arabidopsis thaliana deficient in a chloroplastic starchhydrolyzing<br />

enzyme. The Plant Journal 15, 357–365. doi: 10.1046/<br />

j.1365-313X.1998.00213.x<br />

Zeeman SC, ap Rees T (1999) Changes in carbohydrate metabolism<br />

<strong>and</strong> assimilate partitioning in starch-excess mutants of Arabidopsis.<br />

Plant, Cell & Environment 22, 1445–1453. doi: 10.1046/j.1365-<br />

3040.1999.00503.x<br />

Zeeman SC, Pilling E, Tiessen A, Kato L, Donald AM, Smith AM (2002)<br />

Starch synthesis in Arabidopsis; granule synthesis, composition <strong>and</strong><br />

structure. Plant Physiology 129, 516–529. doi: 10.1104/pp.003756<br />

Zeeman SC, Thorneycroft D, Schupp N, Chapple A, Weck M, Dunstan H,<br />

Haldimann P, Bechtold N, Smith AM, Smith SM (2004) The role of<br />

plastidial α-glucan phosphorylase in starch degradation <strong>and</strong> tolerance<br />

of abiotic stress in Arabidopsis leaves. Plant Physiology 135, 849–858.<br />

doi: 10.1104/pp.103.032631<br />

Zeeman SC, Smith SM, Smith AM (2007) The diurnal metabolism of leaf<br />

starch. The Biochemical Journal 401, 13–28. doi: 10.1042/BJ20061393<br />

Manuscript received 28 November 2006, accepted 7 February 2007

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!