27.07.2014 Views

Amino Acid Transport

Amino Acid Transport

Amino Acid Transport

SHOW MORE
SHOW LESS
  • No tags were found...

Create successful ePaper yourself

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

US ~ epar~ent of Agriculture-~~S and ~epartment of Plant ~iology,<br />

~niversi~ of ~llinois at Urb~na-Champai~n, Urbana, Illinois<br />

*<br />

of the uni~ue features of lants as multicellular organisms is their ability to synthes<br />

from simple inorganic salts and the free ener~y made<br />

tr~sduction reactions of photosyn~esis.<br />

no acids. ~lthoug~ every pla<br />

capacity to assi~late inorganic ni<br />

ly restricted to cells in mature leave<br />

no acids from sites of pri<br />

trogen. I ~ aminoacids ~ s ~ e ~<br />

esis, they are also the precursors of every ~~ogen<br />

luding nucleic acids, gro~th regulators, ~hotosynarray<br />

of other essential compounds. In m ~ y<br />

of ~~ogen me~bolis~ in the plant and underd<br />

bet~een organs is a ~ndamental ~uestio~ in<br />

ids in plants must neces<br />

from the soil solution by the<br />

h onium ions<br />

ssl~lated in the root, or


3 ush<br />

general patterns have been identified (Pate 1980, 1983; Andrews 1986). Temperate perennials<br />

and legumes frequently carry out primary assimilation in the roots, whereas<br />

tropical and subtropical species strongly favor leaf assimilation. Temperate annuals fall<br />

into both categories. Significantly, if high concentrations of nitrate are available in the<br />

soil solution (>1 mM), many species carry out primary assimilation in the leaf. Even<br />

symbiotic nitrogen assimilation is suppressed in the presence of high levels of nitrate<br />

(Harper 1994; Pate 1980).<br />

Once assimilated into amino acids, there are several pathways for the long-distance<br />

transport of reduced nitrogen (Fig. 1). For those plants and conditions in which primary<br />

assimilation occurs in the leaf, amino acids are transported to the heterotrophic tissues<br />

by the phloem-translocation stream. In species that use an apoplastic phloem-loading<br />

mechanism (Bush 1992; Turgeon 1996), this requires the release of amino acids from<br />

mesophyll cells into the apoplastic space, with subsequent loading into the phloem. A<br />

few amino acids (generally a combination of two or more of the following: glutamate,<br />

aspartate, glutamine, asparagine, alanine, and serine) are more concentrated in the<br />

phloem-translocation stream than in the cytoplasm of the mesophyll (Riens et al. 1991;<br />

Winter et al. 1992; Lohaus et al. 1994; Lam et al. 1995). These amino acids must be<br />

actively transported into the phloem (Bush 1993a). In contrast, the measured concentrations<br />

of other amino acids in the phloem suggests they are at or below ~ermodyn~c<br />

equilibrium with the mesophyll and passive transport activity has been suggested (Riens<br />

et al. 199 1; Winter et al. 1992; Lohaus et al. 1994). However, it is important to recognize<br />

that the phloem represents a dynamic pool of amino acids that is rapidly turning over as<br />

pressure-driven mass flow continuously removes the loaded metabolites. Thus, although<br />

snap-shot determinations of amino acid concentrations may implicate passive transport<br />

mechanisms for some amino acids, the kinetics of amino acid turnover suggests that high<br />

rates of transport activity must occur, and passive transport activity, for which the rate<br />

of transport is directly proportional to the concentration difference, may not be rapid<br />

enough to maintain the phloem pools that are quickly flushed away. Given the dynamics<br />

of rapid exchange at the phloem interface, it is not surprising that several high-affinity,<br />

proton-coupled amino acid transporters have been described in plasma membrane vesicles<br />

isolated from mature leaf tissue (Li and Bush 1990, 1991, 1992; Williams et al.<br />

1990, 1992, 1996; Weston et al. 1995). These active transporters can maintain high rates<br />

of flux, even against substantial concentration differences. Nevertheless, evidence for<br />

low levels of facilitated diffusion is also present in membrane vesicles isolated from leaf<br />

tissue (Williams et al. 1990; D. R. Bush unpublished data). Although the participation<br />

of facilitated transporters in phloem loading has yet to be demonstrated, such porters are<br />

likely candidates to play an important role in releasing amino acids from the surrounding<br />

mesophyll cells.<br />

When nitrate or ammonium ions are assimilated in the root, amino acids are transported<br />

to mature leaf tissue in the xylem by the transpiration stream (Pate 1980, 1983).<br />

Because the nitrogen requirement of mature leaves is low, a substantial portion of the<br />

amino acids arriving with the transpiration stream are cycled into the phloem for redistribution<br />

to heterotrophic sinks (see Fig. 1). Cycling amino acids from the xylem to the<br />

phloem in mature leaf tissue is also a fundamental process in symbiotic nitrogen fixation.<br />

For example, atmospheric nitrogen assimilated in the Rhizobium-legume symbiosis is<br />

exported from the root in the transpiration stream as amide amino acids (examples include,<br />

clover and alfalfa) or ureides (mung bean and soybean; Schubert 1986; Winkler<br />

et al, 1988). Even though some of the amide amino acids may be removed from the


<strong>Amino</strong> <strong>Acid</strong> Import<br />

Developing leaves,<br />

<strong>Amino</strong> <strong>Acid</strong> Import<br />

<strong>Amino</strong> <strong>Acid</strong> Cycling<br />

(xylem to phloem)<br />

<strong>Amino</strong> <strong>Acid</strong> Import<br />

<strong>Amino</strong> <strong>Acid</strong> Export<br />

Primary ~ssimilation<br />

<strong>Amino</strong> <strong>Acid</strong> Cycling<br />

(phloem to xylem)<br />

/<br />

<strong>Amino</strong> <strong>Acid</strong> Import<br />

- - = Xylem<br />

Phloem<br />

Figure 1 Schematic presentation of amino acid transport pathways in the plant: <strong>Amino</strong> acids<br />

synthesized in the root from primary assimilation or symbiotic relations are transported to mature<br />

leaf tissue in the xylem (----). <strong>Amino</strong> acids synthesized in the leaf, or arriving from the root in the<br />

xylem, are transported out of the leaf in the phloem (-) to satisfy the needs of heterotrophic<br />

sids. Import-dependent tissues include developing leaves, roots, cortical cells in the stem, seed<br />

and fruits, and apical meristems.<br />

transpiration stream by stem tissues (McNeil et al. 1979; Pate 1983), much of that nitrogen<br />

arrives in mature leaf tissue, from whichit is redistributed by the phloem. The<br />

ureides also arrive in the leaf, but instead of being redistributed in the phloem, they are<br />

catabolized to C02 and (Winkler et al. 1988). Released ammonia is incorporated<br />

into amino acids that are used in leaf metabolism or redistributed to heterotrophic tissues<br />

by the phloem.<br />

Nitrogen cycling between the vascular systems is not restricted to the leaf tissue.<br />

Several lines of evidence show that amino acids can also move from the phloem to the


xylem in the root (Pate 1983; Cooper and Clarkson 1989; Schurr and Schulze 1995).<br />

Thus, cycling amino acids between the two vascular systems in mature leaf and root<br />

tissue (see Fig. 1) permits dynamic adjustments of carbon and nitrogen metabolism in<br />

response to developmental queues or environmental change. For example, the metabolic<br />

needs of the mature root system emphasizes catabolic reactions associated with maintenance<br />

reactions versus the anabolic reactions that drive early growth. Therefore, excess<br />

amino acids transported to the root under these conditions can be redirected to the shoot<br />

for subsequent incorporation into developing seed. Significantly, excess amino acid nitrogen<br />

arriving at the root may also represent a dynamic signal that triggers the downregulation<br />

of nitrate assimilation (Cooper and Clarkson 1989; Crawford 1995; Padgett<br />

and Leonard 1996).<br />

In addition to its role in primary assimilation, amino acid transport is a key process<br />

in leaf senescence and seed germination. In rice, for example, more than 60% of the<br />

amino nitrogen delivered to developing leaves and ears is derived from amino acids<br />

retrieved from senescing older leaves (Mae et al. 1983, 1985; Feller and Fischer 1994).<br />

Likewise, amino acids released from storage proteins during seed ge~nation are the<br />

principal nitrogen source in growing seedlings (Schobert and Komor 1989). Taken together,<br />

it seems clear that amino acids are the currency of nitrogen exchange in the plant.<br />

<strong>Amino</strong> acids are transported into plant cell by proton-coupled symporters that link translocation<br />

across the plasma membrane to the proton-motive force generated by the €I?-<br />

pumping ATPase (Fig. 2; Bush 1993a). These are widely expressed carriers that appear<br />

to be the primary pathways for amino acid transport into plant cells. However, there is<br />

some evidence for facilitated transporters that mediate passive amino acid transport, although<br />

little biochemical and no molecular characterization of such porters has been<br />

published.<br />

The amino acid symporters have been investigated in recent years using purified<br />

membrane vesicles and imposed proton electrochemical potential differences (Bush<br />

1993a). Purified membrane vesicles are a very useful experimental system for studying<br />

membrane transporters. Although membrane transport activity can be examined with<br />

intact cells, there are many complications associated with metabolism and intracellular<br />

comp~mentation that limit experimental interpretation. Additionally, it is difficult to<br />

dissect the bioenergetics of a transport system in intact tissues because of complex interactions<br />

among the primary pumps, ion channels, and other unrelated porters. Isolated<br />

membrane vesicles, on the other hand, allow one to focus on specific transport processes<br />

while minimizing problems associated with the living cell. The major attributes of the<br />

membrane vesicle approach include (1) unequivocal identification of membrane location<br />

using purified vesicles, (2) elimination of posttransport metabolism and compartmentation,<br />

and (3) the ability to control both intra- and extravesiculq solution composition<br />

(Bush 1992). The ability to control solution co~position on both sides of the membrane<br />

is particularly useful for investigating the bioenergetics of the transport process. With<br />

this approach, the basic transport properties of the plant amino acid symporters have<br />

been described (Bush 1993a).<br />

Plant amino acid symporters are electrogenic transporters. Electrogenicity was


Model of amino acid symporter-mediated transport into plant cells: The H+-ATPase<br />

generates a substantial proton electrochemical potential difference (both ApH and AY) that can<br />

drive transport reactions three or four orders of magnitude away from equilibrium. A diagrammatic<br />

representation of the 11 transmembrane domains (TMDs) of NAT2/AAPl is also presented.<br />

demonstrated in isolated membrane vesicles by examining the effect of membrane electrical<br />

potential (AY) on ApH-dependent transport activity. Potassium gradients and valinomycin<br />

were used to manipulate AY. In the absence of compensating charge flow, a low<br />

rate of amino acid flux was observed. When the AY was clamped at zero, transport was<br />

stimulated fourfold, and when a negative AY was imposed, transport was stimulated<br />

sixfold. These data are consistent with an electrogenic transport system that results in<br />

primary charge separation as proton cotransport drives substrate accumulation (Li and<br />

Bush 1990). Similar results were reported for this porter expressed in Xenopus oocytes<br />

(Boorer et al. 1996).<br />

The stoichiometry of H"-amino acid cotransport was recently measured using electrophysiologic~<br />

methods to measure conductance through another plant amino acid symporter<br />

(AAP5) expressed in Xenopus oocytes (Boorer and Fischer 1997). Boorer and<br />

Fischer (1997) used direct measurements of [3H]amino acid uptake and simultaneous<br />

measurements of inward current to show that the stoich~o~etry of the amino acid symporters<br />

is one w' per one amino acid. This held true for the basic, neutral, and acidic<br />

amino acids examined. Because the acidic amino acids also carried one positive charge,<br />

these results suggest they are transported in their neutral forms.<br />

The number of symporters that mediate amino acid transport into higher-plant cells<br />

has yet to be determined. <strong>Transport</strong> competition experiments with isolated membr~e<br />

vesicles provided evidence of at least four amino acid symporters: an acidic amino acid<br />

symporter, a basic amino acid symporter, and two symporters for the neutral amino acids<br />

(Li and Bush 1990, 1991). The neutral amino acid porters were resolved in the competition<br />

experiments because of their differential affinity for isoleucine, valine, threonine,<br />

and proline. A more thorough examination of the kinetics of inter-amino acid transport


competition among the neutral amino acids confirmed the initial conclusion. Li and Bush<br />

(1991) hypothesized that branching at the fharbon in isoleucine, valine, and threonine<br />

results in steric interactions that block their access to one of the neutral porters. More<br />

recently, several amino acid symporters have been cloned, and we now know that many<br />

amino acid symporters are found in the plant (see later discussion).<br />

The ability of various amino acid analogues to compete for transport into purified<br />

membrane vesicles was used to identify molecular determinants that are important for<br />

substrate recognition by the neutral amino acid symporters (Li and Bush 1992). D-ISOmers<br />

of alanine and isoleucine were not effective transport antagonists of the L-isomers,<br />

thereby implicating stereospecificity and suggesting that positional relations between the<br />

a-amino and carboxyl groups is an important parameter in substrate recognition. This<br />

conclusion was supported by the observation that p-alanine and analogues with methylation<br />

at the a-carbon, at the carboxyl group, or at the a-amino group, were not effective<br />

transport antagonists. In contrast, analogues with various substitutions at the distal end<br />

of the amino acid were less potent antagonists. More recently, el~trophysiological investigations<br />

of AAPS expressed in Xenopus oocytes also identified the a-amino and carboxyl<br />

groups, as well as the P-carbon, as important dete~inants in defining substrate<br />

specificity (Boorer and Fischer 1997). Thus, the binding site for the carboxyl end of the<br />

amino acid is a well-defined space that is characterized by compact, asymmetric positional<br />

relations and specific ligand interactions (Li and Bush 1992).<br />

The amino acid symporters are inhibited by carbonyl cyanide ~-chlorophenyl hydrazone<br />

(CCCP) and diethylpyrocarbonate (DEPC; Bush and Langston-~nkefer 1988; Li<br />

and Bush 1990). CCCP is a protonophore that inhibits amino acid transport by dissipating<br />

the proton-motive force that drives the symport reaction. DEPC is a small molecule that<br />

forms covalent bonds with the imidazole ring of histidine residues. The time-dependent<br />

inactivation of alanine transport in the presence and absence of amino acids showed that<br />

substrate binding protects the symporter from DEPC inactivation (Fig. 3). This observation<br />

is consistent with DEPC binding at (or at least confo~ationally linked to) the<br />

substrate-binding site of the carrier (Bush and Li 1991; see Bush 1993b, for discussion).<br />

Since histidine residues frequently participate in protonation reactions, this observation<br />

implicates a histidine residue in the reaction mechanism.<br />

Initial efforts to identify the amino acid syrnporters by using standard biochemical strategieswerelargelyunsuccessful.Althoughthiswasoftenquitefrustrating,<br />

it was not<br />

too surprising, given that these are very low-abundance, integral membrane prot6ins.<br />

Fortunately, a novel molecular approach was successfully used to clone several plant<br />

amino acid symporters and biochemical analysis has moved forward using these clones<br />

and a variety of expression systems.<br />

The first plant amino acid symporter cloned was identified using ~nctional complementation<br />

of yeast amino acid transport mutants (Frommer et al. 1993; Wsu et ai.<br />

1993). ~acc~aro~~ces cerevisiae strains that are auxotrophic and transport-deficient for<br />

a specific amino acid were tr~sfo~ed with plant cDNA libraries that are constructed<br />

in yeast expression vectors. Those transformants that successfully expressed an appropriate<br />

plant amino acid symporter were easily identified by screening for growth on low<br />

concentrations of the limiting amino acid. Because these yeast strains are both auxotrophic<br />

and transport-deficient, a simple secondary screen in the absence of the limiting


80<br />

40<br />

8 20<br />

0<br />

0.0 2.0 4.0 6.0<br />

ure 3 Time- and concentration-dependent inactivation of proton-coupled alanine transport<br />

into purified plasma membrane vesicles (PMVs) isolated from sugar beet leaf tissue: PMVs were<br />

pretreated with DEPC plus or minus alanine for the indicated times, at which point the reaction<br />

was stopped. The vesicles were pelleted and resuspended in inhibitor-~ee buffer, and transport<br />

activity was measured as described previously (Li and Bush 1990).<br />

amino acid was used to show that growth was dependent on amino acid transport, rather<br />

than complementation of the biosynthetic pathway. This is a powerful approach because<br />

it allows one to clone transporter cDNAs without protein sequence information, and<br />

because it provides a useful expression system for subsequent investigations of the biochemical<br />

properties of the trans~ort protein (Bush 1993a; Tanner and Casper 1996). The<br />

screens used for functional co~ple~entation s~ould also id en ti^ high-affinity facilitated<br />

diffusion transpo~ers. in spite of numerous screens of different expression<br />

libraries in a variety of ast strains, no examples of facilitated transporte~ have<br />

been reported.<br />

The cDNA clone of the first amino acid symporter identi~ed with functional complementation<br />

encoded a protein containing 486 amino acids with a calculated molecul~<br />

mass (MJ of 52.9 kDa (Prommer et al. 1993; Hsu et al. 1993). Hy~opathy analysis of<br />

the deduced amino acid sequence suggests this is a typical hydrophobic membrane protein<br />

with 10 to 12 membrane-spanning domains and three sites of potential N-linked<br />

glycosylation. The transport kinetics, substrate specificity, and inhibitor sensitivity of the<br />

heterologously expressed carrier matched those described for one of the neutral amino<br />

acid symporters ch~acterized in purified membrane vesicles (Li and ush 1991). Hsu et<br />

al. (1993) designated this gene ~A~ for neutral amino acid transporter TI, and Prommer<br />

et al. (1993) termed it ~~~. No homologues were identi~ed in the databases, suggesting<br />

that this plant amino acid sympo~er was the first example of a new class of tr~s~rters.


Fromer’s group identified five additional amino acid transporters (AAP2-AAP6)<br />

that are closely related to NAT2/AAPl (Kwart et al, 1993; Fischer et al. 1995; Rentsch<br />

et al. 1996). They are similar in size and predicted topology, and sequence homologies<br />

range from 70 to 95% amino acid si~larity, thus providing good evidence that these<br />

clones represent a family of amino acid translocators (Fischer et al. 1995; Rentsch et al.<br />

1996). As might be expected for a gene family within a given organism, these porters are<br />

differentiated by a combination of unique expression patterns and substrate specificities.<br />

Individually, the AAP transporters have broad substrate specificity, although each exhibits<br />

some preference (lower Km or higher Yma) for certain classes of amino acids. For<br />

example, AAP4 is particularly active for valine and proline, whereas AAP5 is an effective<br />

lysine transporter (Fischer et al, 1995). Within the family, however., there appears to<br />

be some redundancy, for AAP1, AAP2, AAP4, and Asup6 share similar patterns of substrate<br />

specificity, whereas AM3 and AN5 are similar (Fischer et al. 1995; Rentsch et<br />

al. 1996).<br />

Although similar substrate specificity could lead to widespread functional redundancies<br />

between the AAPs., differential expression patterns appear to separate the various<br />

AAI? gene products from one another. For example, AAPS and AAPlNAT2 are widely<br />

expressed in all tissues, whereas AAP3 expression is restricted to the root (Fischer et al.<br />

1995). Moreover, even when related AAPs are expressed in the same organ, they can be<br />

functionally isolated by restricting expression to different cell types within that organ.<br />

Four other classes of plant amino acid transporters have been cloned that are not<br />

members of the AAP gene family. Two of them are basic amino acid ~anspo~ers that<br />

were also cloned by functional complementation: UTI encodes a lysine and arginine<br />

transpo~er that has 533 amino acids and 14 putative transmembrane domains (Frommer<br />

et al. 1995). It is a single-copy gene in ~r~bi~o~sis and sequence alignments suggest it<br />

is closely related to a human basic amino acid transporter (Yoshimoto et al. 1991; Christensen<br />

1992). The second basic amino acid transporter cloned is LHTl (Chen and Bush,<br />

1997). It is a lysine and histidine transporter that contains 446 amino acids and 9-10<br />

transmembrane domains. In contrast to AATl , LHTl does not transport arginine. RNA<br />

gel blot analysis and whole-mount in situ hybridization indicate that LHTl is most<br />

strongly expressed in pollen, siliques, and on the root surface. AATl is most strongly<br />

expressed in vascular tissue and flowers (Frornmer et al. 1995). The expression patterns<br />

of LHTl and UT1 complement one another, suggesting ~HTl may be involved in<br />

amino acid acquisition, whereas AATl may contribute to long-distance transpo~. In<br />

addition to these basic amino acid transporters, an aromatic amino acid transporter has<br />

also been recently characterized.<br />

An aromatic amino acid transporter was identified using functional expression of<br />

an EST cDNA that exhibited low levels of sequence homology to NAT2/AAPl (L. Chen<br />

and D. R. Bush, unpublished data). The cDNA is 1.6 kb, with an open-reading frame<br />

that codes for a protein with 432 amino acids and a calculated M, of 50 ma. Hy~opathy<br />

analysis of the deduced amino acid sequence suggests this is an integral membrane protein<br />

with 11 putative transmembrane domains. This protein exhibits sequence si~l~ty<br />

with AUX1 (a putative auxin transporter; Bennett et al. 1996) and to tryptophan and<br />

tyrosine translocators in E. coli (Wookey and Pittard 1988; Sarsero et al. 1991; Heatwole<br />

and Somerville 1991). RNA gel blot analysis has shown that this transporter is expressed<br />

at very low levels and that it is in all tissues. Whole-mount in situ hybridization further<br />

localized ART1 expression on the surface of roots and leaves and in young seedlings


and in stomata. Overall, ART1 belongs to<br />

favors aromatic amino acids and, perhaps,<br />

Chen and D. R. Bush, unpublished data).<br />

Two proline transporters have also<br />

a new class of amino acid transporter that<br />

to other aromatic compounds in plants (L.<br />

been recently cloned, ProTl and Pro72<br />

(Rentsch et al. 1996). These transporters were identified by complementing a Shr3- yeast<br />

strain that is unable to target yeast amino acid transporters to the plasma membrane.<br />

Significantly, the plant amino acid transporters are not affected by this processing mutation;<br />

therefore, several amino acid transporters (many of which were previously identified)<br />

complemented this yeast strain. A survey of different amino acids for evidence of<br />

transport through these porters showed that they are proline translocators. However, the<br />

rate of proline transport in yeast cells expressing these porters is 300 times slower than<br />

proline transport by AAP6 expressed in the same cell line. It is not known if this is a<br />

result of lower expression levels, or if it is an intrinsic property of these transporters.<br />

ProTl and ProT2 are closely related genes that encode typical membrane proteins that<br />

contain 442 and 439 amino acids, and have 10 putative transmembrane domains. Both<br />

genes are widely expressed in Arabidopsis, with ProTl strongest in roots, flowers and<br />

stems. Significantly, Pro22 expression was enhanced under conditions of water or salt<br />

stress, which is consistent with proline’s putative role in water and salt tolerance (Delauney<br />

and Verma 1993).<br />

Molecular descriptions of amino acid transport have included the isolation of several<br />

transport mutants in Arabidupsis. Two mutants that lack a low-affinity basic amino<br />

acid transport activity were isolated based on their resistance to lysine plus threonine or<br />

to ~-2-~inoethyl-~-cysteine, respectively (Heremans et al. 1997). The specificity of the<br />

lost transport activities were resolved based on apparent K, and sensitivity to transport<br />

inhibitors. Although both mutations have been mapped to chromosome 1, it is not clear<br />

if they are allelic. In a similar approach, the raz1 mutant of Arabidopsis was selected<br />

based on its resistance to the toxic proline analogue, azetidine-2-carboxylic acid (Verbruggen<br />

et al. 1996). Likewise, we have isolated two T-DNA-tagged lines that are resistant<br />

to high concentrations of valine in their growth medium (Chen and Bush 1994).<br />

Exogenous valine is a powerful plant growth inhibitor because it is a negative regulator<br />

of acetohy~ozyacid synthase (also known as acetolactate synthase), an enzyme in the<br />

valine biosynthetic pathway. Once down-regulated in the presence of excess valine,<br />

growth is inhibited because the cells starve for other amino acids, the precursors of which<br />

also pass through acetohy~oxyacid synthase. We were able to show that our lines are<br />

transport mutants versus regulation mutants because they are also resistant to azaserine,<br />

a toxic amino acid analogue that targets other enzymes in the cell. Thus, resistance to<br />

azaserine is consistent with loss of transport activity. We are currently attempting to<br />

identify the disrupted gene(s).<br />

One of the unexpected observations emerging from recent molecular descriptions<br />

of plant transporter genes is that there are large families of related porters that function<br />

in assimilate partitioning (Bush et al. 1996). Examples include the AHA family of protonpumping<br />

ATPases (Harper et al. 1990), the major facilitator superfamily of sugar transporters<br />

(Griffith et al. 1992; Bush et al. 1996), and the amino acid transporter families<br />

reviewed here. Although it is easy to imagine how these porters fulfill complementary<br />

functions in different organs and cells, it is clear that a major challenge in this field is<br />

to determine the unique contributions of the various transporters and, ultimately, to integrate<br />

their combined activities across the plant as a multicellular organism.


Given the central role played by plant amino acid symporters in nitrogen p~itioning<br />

into cells and between organs, and the complexity presented by the many transporters<br />

cloned to date, it is necessw to obtain basic information about their structure and regulation<br />

as the foundation for understanding their contributions to plant growth. Unfortunately,<br />

these kinds of investigations have only recently become possible with the cloning<br />

of the various transporters; therefore, very little has been published in this area. However,<br />

several laboratories have initiated projects investigating st~cture and function, and we<br />

anticipate significant progress in the near future.<br />

itiated a detailed investigation of N<br />

transporter family. We chose NAT<br />

sed, it transports many amino acid e co~on in the<br />

phloem-translocation stream, and because at least one member of this gene fdly appears<br />

to be expressed in every tissue of the plant. We are using an integrated approach<br />

with both biochemical and molecular strategies to learn more about the structure and<br />

function of this protein.<br />

The initial focus of our §tructural analysis has been establishing the membrane<br />

topology of ~ A This ~ is an , important question because our initial analysis suggested<br />

it does not contain 12 membrane-spanning helices as is commonly described for other<br />

metabolite translocators. We started by engineering a c-myc epitope on either the NH2-<br />

or COOH-termini of the expressed protein. We then used in vitro trans~ation, partial<br />

digestion with proteinase , and immuno~recipitation to determine the location of the<br />

NH2- and COOH-termini and to identify a family of oriented peptide fragments that we<br />

resolved on sodium dodecyl sulfate-polyacryldde gel electrophoresis (SDS-PAGE).<br />

These results showed that the ~H2-terminus is located inside the cell and the COOHterminus<br />

is outside. In addition, the lengths of the peptide fragments recovered after<br />

partial proteolysis allowed us to predict the location of protease-accessible cleavage sites<br />

between putative ~ansmembrane domains. We independently identified the location of<br />

the NH2- and COOH-termini using immuno~uorescence microscopy of NAT2 expressed<br />

in COS-1 cells. From the combined data, we proposed an 11 ~ansmembrane domain<br />

model, with the NH2-terminus in the cytoplasm and CO~H-terminus facing outside of<br />

cell (see Fig. 2; Chang and Bush 1997). This model of protein topology anchors our<br />

co~plement~ investigations of porter structure and function using site-directed and<br />

random mutagenesis.<br />

Previous biochemical investigations of ~E~C-dependent inactivation of the amino<br />

acid sympo~ers implicated a histidine residue in the reaction ~echanism (Li and Bush<br />

1990; see Fig. 3). An inspection of the ~ A gene family ~ identi~ed / two ~ histidine ~<br />

is-337, that are conserved in every transporter. Therefore, we used<br />

site-directed mutagenesis to change se residues and then we determined the effect of<br />

those changes on transport activity 0th residues are essential because the modified<br />

proteins were no longer active, owever, the change in is-337 destabiliz~ the protein,<br />

thereby complicating our inte~re~tion of that result. Several amino acid substitutions of<br />

His-47, on the other hand, blocked transport activity in the stable protein, thus demonstrating<br />

the significa~ce of this residue in the transport reaction (L. Chen and D. R. Bush,<br />

un~ublished data).<br />

A second approach we have employed to identify impo~nt &no acid residues<br />

and protein domains uses random mutagenesis and highly selective screens. The advan-


tage of this approach is that it does not require prior biochemical knowledge to identify<br />

residues for potential modification. ~ A ~ was randomly / ~ mutagenized ~ l in a mutator<br />

strain of E. coli and then thousands of modified versions of ~ A were trans- ~ /<br />

formed back into the yeast transport mutants. The yeast were then screened under new<br />

growth conditions that wild-type ~ A ~ does / not permit ~ ~ growth. l For example, we<br />

identified cells that acquired the ability to grow in the presence of transport competitors<br />

and others that grew on very low concentrations of histidine. This kind of positive selection<br />

identified mutant forms of the porter that exhibit new transport properties while<br />

eliminating those that simply lost transport activity. Although loss of function mutants<br />

might be revealing, far too many would inhibit transport activity for unspecific reasons,<br />

such as premature te~ination of the peptide. Thus far, we have identified important<br />

residues in TMDs 1,5,6,8, and 11. ~reliminary analysis of these results suggests TMDs<br />

1, 5, 6, and 8 may be involved in defining the substrate translocation pathway through<br />

this transporter (L. Chen and D. R. Bush, unpublished data).<br />

The amino acid symporters described here are excellent candidates for using biotechnological<br />

methods to modify the nutritional value of harvested tissues. For example, many<br />

cereals are poor sources of protein for animals because they are low t~ptophan, in lysine,<br />

and threonine (Bright and Shewry 1983; z and Larkins 1991 ; Galili 1995), and legumes<br />

are low in cysteine and methionine (Shewry et al. 1995). One strategy for modifying<br />

the nutrition^ value of these deficient crops would be to use targeted expression of<br />

high-affinity amino acid transporters to enhance the content of specific amino acids in<br />

harvested seed. For instance, targeted expression MTl in the phloem of mature leaf<br />

tissue may increase the amount of lysine and histidine transported per unit carbon. This<br />

would increase the lysine and histidine content in developing seeds. The success of this<br />

approach is dependent on several unknown variables, such as reciprocal increases in<br />

mesophyll synthesis and release as these amino acids are actively transported into the<br />

phloem. There is precedence for increased rates of synthesis under increased demand,<br />

and recent work in Heldt’s laboratory suggests the mesophyll pool may be dynamically<br />

linked to phloem loading (Mens et al. 1991 ; Winter et al. 1992; Lohaus et al. 1994) and,<br />

thus, active loading by a high-affinity transporter may shift the equation to increased<br />

synthesis. Other examples of potential engineering include targeted expression in stem<br />

tissue to capture amino acids passing by in the phloem, combining ove~roduction and<br />

targeted expression, or targeted expression to engineer uptake of novel xenobiotics.<br />

Our understanding of amino acid transport has progressed si~nificantly in recent<br />

years. The challenges that lay ahead focus on defining the coarse and fine-con~ol mechanisms<br />

that regulate that function of these amino acid transporters. This knowledge will<br />

be the fo~ndation of our understanding of resource allocation across the plant as a multicellular<br />

organism, and for developing novel strategies for improving crop yields and<br />

nu~tion~ value.<br />

Andrews, M. (1986). The p~itioning of nitrate assi~ilation between root and shoot of higher<br />

plants. Plant Cell Environ., 9: 511-519.<br />

Bennett, M. J., Marchant, A,, Green, . G., May, S. T., Ward, S. P., Millner, P. A., Walker, A,


R., Schulz,B.,andFeldmann,K.A. (1996). Arabidopsis AUX1 gene:Apermease-like<br />

regulator of root gravitropism. Science, 273: 948-950.<br />

Boorer, IC. J., Frommer, W. B., Bush, D. R., Kreman, M., Loo, D. D. F., and Wright, E. M. (1996).<br />

Kinetics and specificity of a €%+/amino acid transporter from Arabidopsis thaliana. J. Biol.<br />

Chem., 271: 2213-2220.<br />

Boorer, K. J. and Fischer, vir. N. (1997). Specificity and stoichiometry of the Arabidopsis H+/<br />

amino acid transporter AAP5. J. Biol. Chem., 272: 13040-13046.<br />

Bright, S. W. J. and Shewry, P, R. (1983). Improvement of protein quality in cereals. CRC Crit.<br />

Rev. Plant Sci., 1: 49-93,<br />

Bush, D. R. (1993a). Proton-coupled sugar and amino acid transporters in plants. Annu. Rev. Plant<br />

Physiol. Plant Mol. Biol., 44: 513-542.<br />

Bush,D. R. (1993b). Inhibitors of the proton-sucrose symport. Arch. Biochem. Biophys., 307<br />

355-360.<br />

Bush, D. R. (1992). The proton-sucrose symport. Photosynth. Res., 32: 155-165.<br />

Bush, D. R., Chiou, T. J., and Chen, L. (1996). Molecular analysis of plant sugar and amino acid<br />

transporters. J. Exp. Bot., 47 1205-1210.<br />

Bush, D. R. and Li, 2;. C. (1991). Froton-coupled sucrose and amino acid transport across the plant<br />

plasma membrane. Recent Advances in Phloem <strong>Transport</strong> and Assimilate Compa~mentation.<br />

(J. L. Bonnemain, S. Delrot, J. Dainty, and W. J. Lucas, eds.), West Publications, Paris,<br />

pp.148-153.<br />

Bush, D. R. and Langston-Unkefer, P. J. (1988). <strong>Amino</strong> acid transport into membrane vesicles<br />

isolated from zucchini. Plant Physiol., 88: 487-490.<br />

Chang, H.-C. and Bush, D. R. (1997). Topology of NAT2, a prototypical example of a new family<br />

of amino acid transporters. J. Biol. Chem. 272: 30552-30557.<br />

Chen, L. and Bush D. R. (1997). LHTl, a lysine and histidine specific amino acid transporter in<br />

Arabidopsis. Plant Physiol, 115: 1127-1 134.<br />

Chen, L. and Bush, D. R. (1994). <strong>Amino</strong> acid transport mutants in T-DNA tagged Arabidopsis.<br />

Plant Physiol., 105s: 149.<br />

Christensen, H. N. (1992). A retrovirus uses a cationic amino acid transporter as a cell surface<br />

receptor. Nutr. Rev., 50: 47-48,<br />

Cooper, H. D. and Clarkson, D. T. (1989). Cycling of amino-nitrogen and other nutrients between<br />

shoots and roots in cereals-a possible mechanism integrating shoot and root in the regulation<br />

of nutrient uptake. J. Exp. Bot., 216 753-762.<br />

Crawford, N. M. (1995). Nitrate-nutrient and signal for plant growth. Plant Cell., 7 859-868.<br />

Delauney, A. J. and Verma, D. P. S, (1993). Proline biosynthesis and osrnoreregulation in plants.<br />

Plant J,, 4: 215-223.<br />

Epstein, E. (1972). Mineral ~utrition of Plants: Principles and Perspectives. John Wiley & Sons,<br />

New York.<br />

Feller, U. and Fischer, A. (1994). Nitrogen metabolism in senescing leaves. Crit. Rev. Plant Sci.,<br />

13: 241-273.<br />

Fischer, W. N., Kwart, M., Humel, S., and Frommer, W. 3. (1995). Substrate speci~~ty and expression<br />

profde of amino acid transporters (AAPs) Ar~i~psis,<br />

in J. Biol, Chem., 2Z 16315-16320.<br />

Frommer,W. B., Hummel, S,, andRiesmeier,J.W. (1993).Expressioncloninginyeastofa<br />

cDNA encoding a broad specificity amino acid permease from Arabidopsis thaliana. Proc.<br />

Natl. Acad. Sci. USA, 90: 5944-5948.<br />

Frommer, W, B., Hummel, S., Unseld, M., and Ninneman, 0. (1995). Seed and vascular expression<br />

of a high affinity transporter for cationic amino acids in Arabidopsis. Proc. Natl, Acad. Sci.<br />

USA, 92: 12036-12040.<br />

Galili, G. (1995). Regulation of lysine and threonine synthesis. Plant Cell, 7 899-906.<br />

Griffith, J. K,, Baker, M. E., Rouch, D. A., Page, M. G. P., Skurry, R. A., Paulsen, I. T., Chater,<br />

K. F., Baldwin, S. A,, and Henderson, P. J. F. (1992). Membrane transport proteins: Implications<br />

of sequence comparisons. Curr. Opin Cell Biol., 4: 684695.


Harper, J. E. (1994). Nitrogen metabolism. Physiology and ~ete~ination of Crop Yield, pp. 285-<br />

302. Ed. K. J. Boote, American Society of Agronomy, Madison, WI.<br />

Harper, J. F., Manney, L., DeWitt, N. D., Yoo, M. H., and Sussrnan, M. R. (1990). The Arabidopsis<br />

thal~ana plasma membrane ~ -A~ase multigene family. Genomic sequence and expression<br />

of the third isoform. J. Biol. Chem., 265: 13601-13608.<br />

Heatwole, V. M. and Somerville, R. L. (1991). Cloning, nucleotide sequence, and characterization<br />

of mtr, the structural gene for a tryptophan-specific permease of Escherichia coli K-12, J.<br />

Bacteriol., 173: 108-1 15.<br />

Heremans, B., Borstlap, A. C., and Jacobs, M. (1997). The rltll and raecl mutants of Ara~idopsis<br />

t~liana lack the activity of a basic-amino-acid transporter. Planta, 201: 219-226.<br />

Hsu, L. C., Chiou, T. J., Chen, L., and Bush, D. R. (1993). Cloning a plant amino acid transporter<br />

by functional complementation of a yeast amino acid transport mutant. Proc, Natl, Acad.<br />

Sci. USA, 907441-7445.<br />

Kriz, A. L. and Larkins, B. A. (1991). Biotechnology of seed crops: Genetic engineering of seed<br />

storage proteins. Hort. Sci., 26: 1036-1041.<br />

Kronzucker, H, J., Siddiqi, M. Y., and Glass, A. D. M. (1997). Conifer root discrimination against<br />

soil nitrate and the ecology of forest succession. Nature, 385: 59-61.<br />

Kwart, M., Hirner, B., Hurnmel, S,, and Frommer, W. B. (1993). Differential expression of two<br />

related amino-acid transporters with differing substrate specificity in Arabidopsis thaliana.<br />

Plant J., 4: 993-1002.<br />

Lam, H.-H., Coschigano, K. T., Oliveira, I. C., Melo-Oliveira, R., and Coruzzi, G. M. (1996). The<br />

molecular-genetics of nitrogen assimilation into amino acids in higher plants. Annu. Rev.<br />

Plant Physiol. Plant Mol. Biol., 47: 569-593.<br />

Lam, H.-M., Coschigano, K. T., Schulta, C., Melo-Oliveira, R., Tjaden, G., Oliveira, I., Ngai, N.,<br />

Hsieh, M.-H., and Coruzzi, 6. M. (1995). Use of Arabidopsis mutants and genes to study<br />

amide amino acid biosynthesis. Plant Cell, 7: 887-898.<br />

Li, Z.-C. and Bush, D. R. (1990). ApH-dependent amino acid transport into plasma membrane<br />

vesicles isolated from sugar beet leaves. I. Evidence for ca~er-mediated, electrogenic flux<br />

through multiple transport systems. Plant Physiol., 94: 268-277.<br />

Li, Z.-C. and Bush, D. R. (1991). ApH-dependent amino acid transport into plasma membrane<br />

vesicles isolated from sugar beet (Beta vulgaris L,) leaves. II. Evidence for multiple aliphatic,<br />

neutral amino acid symporters. Plant Physiol., 96 1338-1344.<br />

Li, Z.-C. and Bush, D. R. (1992). Structural determinants in substrate recognition by proton-amino<br />

acid symports in plasma membrane vesicles isolated from sugar beet leaves, Arch. Biochem.<br />

Biophys., 294: 519-526.<br />

Lohaus, G., Burba, M., and Heldt, H. W. (1994). Comparison of the contents of sucrose and amino<br />

acids in the leaves, phloem sap, and taproots of high and low sugar-producing hybrids of<br />

sugar beet. J. Exp. Bot., 45: 1097-1 101.<br />

Mae,T., Makino, A,,andOhira, K. (1983).Changesintheamountsifribulosebisphosphate<br />

carboxylase synthesized and degraded during the life span of rice leaf (Qryza sativa). Plant<br />

Cell Physiol., 21: 1079-1081.<br />

Mae, T., Hoshino, T., and Ohira, K. (1985). Protease activities and loss of nitrogen in the senescing<br />

leaves of field grown rice (Qryza sativa L.), Soil Sci. Plant Nutr., 31: 589-4500.<br />

McNeil, D. L., Atkins, C. A., and Pate, J. S. (1979). Uptake and utilization of xylem borne amino<br />

compounds by shoot organs of a legume. PZant Physiol., 63: 1076-1081.<br />

Padgett, P. E. and Leonard, R. T. (1996). Free amino acid levels and the regulation of nitrate<br />

uptake in maize cell suspension cultures. J. Exp. Bot., 300: 871-883.<br />

Pate, J. S. (1976). <strong>Transport</strong> in symbiotic systems fixing nitrogen. Encycl. Plant Physiol. New Ser.,<br />

2: 278-303.<br />

Pate, J. S. (1980). <strong>Transport</strong> and partitioning of nitrogenous solutes. Annu. Rev. Plant Physiol., 31:<br />

3 13-340.<br />

Pate,J. S. (1983). Distribution of metabolites. Plant Physiology, A Treatise. Vol. 8: Nitrogen


3<br />

Meta~olism (F. C. Stewart and R. G. S. Bidwell, eds.), Academic Press, New York, pp.<br />

335-401.<br />

Rentsch, D., Himer, B., Schmelzer, E., and Frommer, W. B. (1996). Salt stress-induced proline<br />

transporters and salt stress-repressed broad specificity amino acid permeases identified by<br />

suppression of a yeast amino acid pe~ease-targeting mutant. Plant Cell, 8 1437-1446.<br />

Riens, B., Lohaus, G., Heineke, D., and Heldt, H. W. (1991). <strong>Amino</strong> acid and sucrose content<br />

determined in the cytosolic, chloroplastic, and vacuolar comp~ments and in the phloem<br />

sap of spinach leaves. Plant Physiol., 97 227-233.<br />

Sarsero, J. P., Wookey, P. J., Gollnick, P., Yanofsky, e,, and Pittard, A. J. (1991). A new family<br />

of integral membrane proteins involved in transport of aromatic amino acids in ~scherichis<br />

coli. J. Bacteriol,, 173: 3231-3234.<br />

Schobert, C. and Komor, E. (1989). The differential transport of amino acids into the phloem of<br />

Ricinus co~munis L. seedlings as shown by the analysis of sieve-tube sap. Planta, 177:<br />

342-349.<br />

Schubert, K. R. (1986). Products of biological nitrogen fixation in higher plants: Synthesis, transport,<br />

and metabolism. Annu. Rev. Plant Physiol., 37: 539-579.<br />

Schurr, U. and Schulze, E. D. (1995). The concentration of xylem sap constituents in root exudate,<br />

and in sap from intact, transpi~ng castor bean plants. Plant Cell ~nviron., 18 409-420.<br />

Tanner, W. and Caspari, T. (1996). Membrane transport carriers. Annu. Rev. Plant Physiol. Plant<br />

Mol. Biol., 47: 595-626.<br />

Turgeon, R. (1996). Phloem loading and plasmodesmata. Trends Plant Sci., I: 418-423.<br />

Verbruggen, N., Borstlap, A. C., Jacobs, M., Van Montagu, M., and Messens, E. (1996). The razl<br />

mutant of Ara~idopsis thaliana lacks the activity of a high affinity amino acid transporter.<br />

Planta, 200: 247-253.<br />

Weston, K., Hall, J. L., and williams, L. E. (1995). Characte~zation of amino acid transport in<br />

Ricinus communis roots using isolated membrane vesicles. Planta, 196 166-173.<br />

Williams, L. E., Nelson, S, J., and Hall, J, L. (1990). Characte~zation of solute transport in plasma<br />

membrane vesicles isolated from cotyledons of Ricin~s co~munis L. 11: Evidence for a<br />

proton-coupled mechanism for sucrose and amino acid uptake. Planta, 182: 540-545.<br />

Williams, L. E., Nelson, S. J., and Hall, J, L. (1992). Characte~zation of solute/proton cotransport<br />

in plasma membrane vesicles from Ricjn~s cotyledons, and a comparison with other tissues.<br />

Planta, 186 541-550.<br />

Williams, L. E., Bick, J. A., Neelam, A., Weston, K. N., and Hall, J, L, (1996). Biochemical and<br />

molecular characte~zation of sucrose and amino acid carriers in ~icinus communis. J. Exp.<br />

Bot., 47: 1211-1216.<br />

Winkler, R. G., Blevins, D. G., Polacco, J. C., and Randall, D. D. (1988). Ureide catabolism in<br />

nitrogen fixing legumes. Trends Biochem. Sci., 13: 97-100.<br />

Winter, H., Lohaus, G., and Heldt, H. W. (1992). Phloem transport of amino acids in relation to<br />

their cytosolic levels in barley leaves. Plant Physiol., 99: 996-1004.<br />

Wookey, P. J. and Pittard, A. J. (1988). DNA sequence of the gene (tryP) encoding the tyrosinespecific<br />

transport system of ~scherichia coli. J. Bacteriol,, 170 494644949.<br />

Yoshimoto, T,, Yoshimoto, E., and Meruelo, D. (1991). Molecular cloning and characte~zation of<br />

a novel human gene homologous to the murine ecotropic retroviral receptor. Virology, 185:<br />

10-17.


David Rhodes<br />

Purdue Unjversj~, West ~fa~ette, Indiana<br />

~n~ve~si~ of M~sso~~i, Colu~bia, Missouri<br />

1. INT TI<br />

This chapter will consider the central role of amino acid metabolism in abiotic stress<br />

resistance of plants. We will attempt to ~ighlight progress made since the last comprehensive<br />

review of this field by Stewart and Larher (1980). Major emphasis will be placed<br />

on the role of amino acid metabolism in the synthesis of compatible osmolytes. This<br />

class of molecules includes certain amino acids (notably proline), quaternary ammonium<br />

compounds (e.g., glycinebetaine, prolinebetaine, P-alaninebetaine, and choline-0-sulfate),<br />

and the tertiary sulfonium compound 3-dim~thylsulfoniopropionate (DMSP), The<br />

quaternary ammonium compounds and DMSP are derived from amino acid precursors.<br />

These compounds share the property of being uncharged at neutral pH and are of high<br />

solubility in water (Ballantyne and Chamberlin 1994). Moreover, at high concen~ations<br />

they have little or no perturbing effect on macromolecule-solvent interactions (Yancey<br />

et al. 1982; Low 1985; Somero 1986; Timasheff 1993; Yancey 1994). Unlike perturbing<br />

solutes (such as inorganic ions) that readily enter the hydration sphere of proteins, favoring<br />

unfolding, compatible osmolytes tend to be excluded from the hydration sphere of<br />

proteins and stabilize folded protein structures (Low 1985). These compounds are<br />

thought to play a pivotal role in plant cytoplasmic osmotic adjustment in response to<br />

osmotic stresses (Wyn Jones et al. 1977).<br />

The synthesis of quate~ary ammonium and tertiary sulfonium compounds requires<br />

participation of the “activated methyl cycle’’ to provide methyl groups for the onium<br />

moieties of these compounds (Hanson et al. 1995; Bohnert and Jenson 1996). Therefore,<br />

attention is given to the role of sulfur amino acid metabolism, and in particular, the role<br />

of methionine and ~-adenosylmethionine (SAM) in the biosynthesis of these methylated<br />

onium compounds. The central position of SAM in stress metabolism in plants is further<br />

illustrated by briefly considering its role as a precursor of the hormone ethylene (Kende<br />

1983) and of the polyamines spermine and spermidine (Flores et ai. 1989). The role of<br />

sulfur amino acid metabolism in plant stress-resistance mechanisms is further explored<br />

in relation to the p~icipation of cysteine as a precursor of peptides (glutathione and<br />

319


phytochelatins) that chelate heavy metals (Rauser 1990, 1995) and facilitate detoxification<br />

of herbicides (Li et al. 1995; Kreuz et al. 1996) and active oxygen species (Alscher<br />

1989; Foyer et al. 1995; Smirnoff 1995).<br />

The amino acids alanine and y-aminobutyrate (CABA) accumulate markedly in<br />

response to anaerobic stress in plants (Streeter and Thompson 1972); these amino acids<br />

are specifically discussed in relation to anaerobic carbon metabolism, and intracellular<br />

pH regulation. Because it is now recognized that CABA synthesis is catalyzed by a<br />

C~-calmodulin-activated enzyme, glutamate decarboxylase (Snedden et al. 1995), we<br />

will consider the specific role of GfU3A connecting intermediary amino acid metabolism<br />

to perturbations of cytoplasmic Ca" concentrations induced by stress.<br />

Because proline is treated in a separate chapter in this volume (see Chap. 8), and has<br />

been the subject of numerous reviews over the last 20 years (see, e.g., Stewart and Larher<br />

1980; Thompson 1980; Stewart 1981; Hanson and Hitz 1982; Csonka and Baich 1983;<br />

Rhodes 1987; Delauney and Verma 1993; Samaras et al. 1995), including a comprehensive<br />

recent review (Hare and Cress 1997), our treatment of this amino acid will be brief.<br />

Proline accumulation represents a common metabolic responses of higher plants to water<br />

deficits and salinity stress. This highly water-soluble imino acid is accumulated to osmotically<br />

significant levels by leaves of many halophytic higher-plant species grown in saline<br />

environments (Stewart and Lee 1974; Treichel 1975; Briens and Larher 1982), in leaf<br />

tissues and shoot apical meristems of plants experiencing water stress (Barnett and<br />

Naylor 1966; Boggess et al. 1976; Jones et al. 1980), in desiccating pollen (Hong-qi et<br />

al. 1982; Lansac et al. 1996), and in cultured plant cells adapted to water stress (Handa<br />

et al. 1986; Rhodes et al. 1986), or NaCl stress (Katz and Tal 1980; Treichel 1986;<br />

Binzel et al. 1987; Rhodes and Handa 1989; Thomas et al. 1992). In the apical millimeter<br />

of maize roots, proline represents a major solute, reaching concentrations of 120 mM in<br />

roots growing at a water potential of -1.6 MPa (Voetberg and Sharp, 1991). The accurnulated<br />

proline accounts for a significant fraction ("50%) of the osmotic adjustment in this<br />

region (Voetberg and Sharp 1991). Proline accumulation in maize root apical meristems<br />

in response to water deficits involves increased proline deposition to the growing region<br />

(Voetberg and Sharp 1991) and requires abscisic acid (ABA; Ober and Sharp 1994;<br />

Sharp et al. 1994).<br />

At high concentrations, proline protects membranes and proteins against the adverse<br />

effects of inorganic ions and temperature extremes (reviewed in Samaras et al.<br />

1995). Proline may also function as a hydroxyl radical scavenger (Smirnoff and Cumbes<br />

1989). Its synthesis is implicated as a mechanism of alleviating cytoplasmic acidosis,<br />

and it may maintain NADP+/NAPH ratios at values compatible with metabolism (Hare<br />

and Cress, 1997). Rapid catabolism of proline after relief of stress may provide reducing<br />

equivalentsthatsupportmitochondrialoxidativephosphorylationandthegeneration of<br />

ATP for recovery from stress and repair of stress-induced damage (Hare and Cress, 1997).<br />

In vivo-labeling studies with 14C-labeled precursors (Moms et al. 1969; Oaks et al.<br />

1970; Boggess et al. 1976), ["CC]glutamate (Hayser et al. 1989a,b) or ''NH~/''NO~<br />

(Rhodes et al. 1986; Rhodes and Handa 1989) suggest that glutamate is a major precursor


of osmotic stress-induced proline accumulation in plants. Osmotic stress results in an<br />

increase of proline biosynthesis rate (Boggess et al. 1976; Stewart, 1981; Rhodes et al.<br />

1986; Rhodes and Handa 1989). Proline accumulation involves induction of enzymes of<br />

proline biosynthesis (Peng et al. 1996; and see Chap. 8), possibly coupled with a relaxation<br />

of proline feedback-inhibition control of the pathway (Boggess et al. 1976; Stewart<br />

1981), decreased proline oxidation to glutamate (Stewart et al. 1977; Stewart and Boggess<br />

1978; Wuang and Cavalieri 1979; Sells and Koeppe 1981; Elthon and Stewart 1982),<br />

mediated at least partly by down-regulation of proline dehydrogenase (Kiyosue et al.<br />

1996, Peng et al. 1996), decreased utilization of proline in protein synthesis (Boggess<br />

and Stewart 1980; Stewart 1981), and enhanced protein turnover (Fukutoku and Yamada<br />

1984). Water deficits induce dramatic increases in the proline concentration of phloem<br />

sap in alfalfa (Girousse et al. 1996), suggesting that increased deposition of proline at<br />

the root apex in water-stressed plants (e,g., Voetberg and Sharp 1991) could occur partly<br />

by phloem transport of proline (Girousse et al. 1996). A proline transporter gene ProT2<br />

is strongly induced by water and salt stress in Arnbidopsis t~~Ziann (Rentsch et al. 1996).<br />

In bacteria, proline synthesis from glutamate is catalyzed by three enzymes: y-<br />

glutamyl kinase (GK; EC 2.7.2.1 I), y-glutamyl phosphate reductase (GPR; glutamic semialdehyde<br />

dehydrogenase; EC 1.2.1 Al), and A1-pyrroline-5-carboxylate reductase (PSCR;<br />

EC 1.5,1.2), encoded by genes pro& proA, and yroC, respectively (Adams and Frank<br />

1980; Csonka and Baich 1983). In higher plants it is now clear that a similar pathway<br />

operates, with the exception that GK and GPR are embodied in one bifunctional polypeptide,<br />

A1-py~oline-5-carboxylate synthetase (PSCS), encoded by a single gene (Hu et al.<br />

1992; Fig. 1). The prevailing evidence is that both PSCS and PSCR are localized in the<br />

cytoplasm (Hare and Cress, 1997). The GK activity of the bifunctional PSCS is inhibited<br />

by proline (Nu et al. 1992). PSCS gene expression is induced by desiccation, salinity<br />

stress, and abscisic acid (ABA; Hu et al. 1992; Yoshiba et al. 1995; Peng et ai. 1996;<br />

Savourk et al. 1997). P5CR transcripts also increase in abundance in response to osmotic<br />

stress (Delauney and Verma 1990; Williamson and Slocum, 1992; Verbruggen et al.<br />

1993). Savourd et al. (1997) suggest that the expression of the proline biosynthetic genes<br />

are dependent on at least two signal transduction cascades: one triggered by exogenously<br />

applied ABA in the absence of stress, and the other triggered by cold and osmotic stress<br />

independently of exogenously applied ,ABA.<br />

Recently, transgenic tobacco plants overexpressing PSCS have been obtained, and<br />

these appear to have increased resistance to both water deficits and salinity stress (Kishor<br />

et al. 1995). However, insufficient data are available to conclude with certainty that<br />

proline accumulation in these transgenic plants contributes to their enhanced stress resistance<br />

by osmotic adjustment, or by some other mechanisms (Sharp et al. 1996). As discussed<br />

by Hare and Cress (1997), the controversial data reported for the water relations<br />

of these transgenic plants emphasize the need to investigate nonosmotic explanations for<br />

the phenotypes observed.<br />

A decrease in proline oxidation rate can contribute to net proline accumulation<br />

during drought and salinity stress. Proline is oxidized to PSC by proline dehydrogenase<br />

(EC 1.5.99.8) and the resulting P5C is further oxidized to glutamate by P5C dehydrogenase<br />

(EC 1.5.1.12; see Fig. 1). Both enzymes are localized in the mitochondrion (reviewed<br />

in Hare and Cress, 1997). A cDNA-encoding proline dehydrogenase (PDH) has<br />

recently been cloned from Arn~i~opsis and is induced by proline under nonstressed conditions,<br />

but strongly repressed in response to osmotic stress (Kiyosue et al. 1996). It<br />

appears that osmotic stress overrides proline induction of proline dehydrogenase gene


OH glutamate +<br />

8<br />

0 0<br />

*.<br />

I<br />

*<br />

*.<br />

NADH<br />

+ H+<br />

NADS<br />

* *<br />

.I<br />

0 .<br />

.*.*<br />

PSCR<br />

H O<br />

proline -<br />

P<br />

proline<br />

FADH2<br />

+ H20<br />

FAD +<br />

c<br />

+ 02<br />

iigure 1 Pathways of synthesis and oxidation of proline in higher plants: GK, y-glutamyl kinase;<br />

GPR, y-glutamylphosphate reductase; PSC, A1-pyrroline-5-~arboxylate; PSCS, A'-pyrroline-S-carboxylate<br />

synthetase; PSCR, A'-pyrroline-5-carboxylate reductase; PDH, proline dehydrogenase;<br />

PSCDH, A1-pyrroline-S-c~boxylate dehydrogenase. The dotted line denotes feedback inhibition of<br />

the GK activity of PSCS by the end product, proline,


expression (Kiyosue et al. 1996). The reciprocal regulation of PSCS and PD<br />

count for the rapid accumulation of proline in response to osmotic stress and its rapid<br />

catabolism after stress relief (Peng et al. 1996). To our knowledge the gene encoding<br />

P5C dehydrogenase has not yet been cloned from higher plants, but the enzyme has<br />

recently been extensively purified and characterized (Forlani et al. 1997).<br />

Hare and Cress (1997) argue that the different subcellular localizations of proline<br />

biosynthesis (cytoplasm) and oxidation (mitochondrion), the NADPH cofactor preference<br />

of the biosynthetic enzymes, and NADH cofactor preference for the proline oxidation<br />

pathway, would enable proline biosynthesis to enhance activity of the cytoplasmic oxidative<br />

pentose phosphate pathway and provide a mechanism of interconversion of the phosphorylated<br />

and nonphosphorylated pools of pyridine nucleotide cofactors. They suggest<br />

that the osmoprotective effects of proline accumulation may be of secondary impo~ance<br />

to the associated metabolic implications of proline synthesis and degradation. Further<br />

investigations of transgenic plants, engineered for altered proline synthesis or catabolism,<br />

clearly need to consider not only the consequences on absolute proline level and the<br />

biophysical effects exerted by elevated cytoplasmic proline concentrations, but also the<br />

fluxes to and from the proline pool (Hare and Cress 1997).<br />

In growing tissue, the proline deposition rate will be equal to the combined rates<br />

of proline synthesis, proline release from protein, and proline import, minus the combined<br />

rates of proline catabolism, proline export, proline utilization in protein synthesis,<br />

and the rate of pool dilution caused by water uptake during growth (Rhodes and Handa<br />

1989; Voetberg and Sharp 1991). The latter, in turn, will be determined by fundamental<br />

plant-water relations of the growing region, including water potential, solute potential,<br />

yield threshold, turgor, and potential difference between the xylem and growing cells<br />

(Nonami et al. 1997), as well as the levels of growth inhibitory metabolites such ABA, as<br />

for which synthesis or compartmentation may respond rapidly to changes in turgor or<br />

metabolism (e.g., intracellular pH; Rhodes 1987). The feedback loops in this system are<br />

clearly complex, with ABA and osmotic signals altering expression of genes encoding<br />

proline biosynthesis enzymes, proline and osmotic signals altering expression of PD<br />

proline feedback inhibiting its own synthesis, and proline itself (if accumulated to sufficiently<br />

high levels) contributing to solute potential and, hence, turgor and growth maintenance.<br />

The full promise of transgenic plants engineered for proline metabolism toward<br />

an understanding of proline’s role(s) in stress resistance will likely not be realized until<br />

attention is given to the growing regions as well as mature tissues (cf., Kishor et al.<br />

1995). The consideration of growing and nongrowing tissues and the fluxes between<br />

them will be essential in testing the intriguing scheme, recently proposed by Hare and<br />

Cress (1997), in which proline and PSC might act as an intercellu1~-signaling system.<br />

In this scheme, proline produced from P5C in an “effector” tissue is transported by the<br />

phloem to a “target” tissue, characterized by a high-energy requirement, where proline<br />

degradation generates reducing equivalents needed to drive the tricarboxylic acid (TCA)<br />

cycle activity. The PSC or glutamate generated may then be translocated back to the<br />

effector tissue where conversion to proline regenerates the NADP’ needed to prime the<br />

oxidative pentose phosphate pathway.<br />

Stress-hypersensitive mutants of higher plants that exhibit disturbed proline metabolism<br />

(e.g., the sosl mutant of Ar~~i~o~sis; Liu and Zhu 1997) can contribute significantly<br />

to the elucidation of thesignalsto which proline accumulation may respond.<br />

However, we urge caution in concluding from such mutants that “Pro accumulation is a<br />

symptom of stress injury rather than an indicator of stress tolerance” (Liu and Zhu 1997).


324 ~ho~es et a!.<br />

111. IN0 ACIDS AS PREC<br />

Y AMMONIUM COMP<br />

OMPATIBLE OSMOLYTES<br />

The role of quaternary ammonium compounds as compatible solutes has been extensively<br />

reviewed (Wyn Jones and Storey 1981 ; Anthoni et al. 199 1 ; Rhodes and Hanson 1993;<br />

Gorham 1995). These zwitterionic compounds include glycinebetaine, choline-0-sulfate,<br />

P-alaninebetaine, prolinebetaine, and hydroxyprolinebetaine. They possess a fully<br />

methyl-substituted nitrogen atom (creating a permanent positive charge on the N moiety)<br />

and a negatively charged carboxyl group (in the case of betaines) or sulfate group (in<br />

the case of choline-0-sulfate). The function of these compounds as osmoprotectants is<br />

illustrated by their stimulation of the growth of Escherichia coli or ~aZ~oneZZa ~ ~ ~ i ~<br />

rium in saline medium (Strom et al. 1983; Hanson et al. 1991, 1994a). In these assays,<br />

glycinebetaine, P-alaninebetaine, prolinebetaine, and choline-0-sulfate at 1-mM concentrations<br />

are more effective than 1 mM of proline (Hanson et al. 1991, 1994a). Choline,<br />

an alcohol without a negatively charged group, is not an osmoprotectant per se. Although<br />

it stimulates the growth of wild-type E. coli in a saline medium, it must be oxidized to<br />

glycinebetaine to function as an osmoprotectant. Thus, betA mutants of E. coli, defective<br />

in choline oxidation, are unable to use choline as an osmoprotectant (Hanson et al. 1991).<br />

Similar to proline, glycinebetaine and structurally related compounds have stabilizing<br />

effects on proteins and membranes (Rhodes and Hanson 1993), and are implicated in<br />

frost protection (Kishitani et al. 1994; Nomura et al. 199Sb).<br />

In higher plants, choline is derived from the amino acid serine, presumably by decarboxylation<br />

of serine to ethanolamine (Fig. 2). The NH?-moiety of ethanolamine is then N-<br />

methylated to (CH&-P- at the level of either free ethanolamine bases, O-phosphorylethanolamine<br />

bases, or 0-phosphatidylethanolamine bases, catalyzed by S-adenosylmethionine<br />

(SAM)-dependent N-methyltransferases (Rhodes and Hanson 1993; Corham<br />

1995; Hanson et .al. 1995). Datko and Mudd (1988a,b) have proposed phospho~lethanolamine<br />

(PE) as a common co~itting step in the synthesis of choline moieties in plants,<br />

implicating ethanolamine kinase (EC 2.7.1.82) as a key enzyme in this pathway (see<br />

Fig. 2).<br />

In glycinebetaine-accumulating chenopods, the main pathway of choline synthesis<br />

appears to be by the phosphoryl-bases: PE, phospho~lmonomethylethanolamine (PMME),<br />

phosphoryldimethylethanolamine (PDME), and phosphorylcholine (PC) (Hanson and<br />

Rhodes 1983; Hanson et al. 1995; see Fig. 2). Choline is then liberated from PC by a<br />

phosphorylcholine phosphatase (Rhodes and Hanson 1993; Hanson et al. 1995; see<br />

Fig. 2).<br />

In grasses, PC is incorporated into phosphatidylcholine from which choline is then<br />

released (Giddings and Hanson 1982; Hitz et al.1.991),presumably by the action of<br />

phospholipase-D (Rhodes and Hanson 1993).<br />

Choline down-regulates its own synthesis at the level of the PMME and PDME N-<br />

methyltransferases (Mudd and Datko 1989a,b). These enzymes are induced by salinity<br />

stress in spinach (Weretilnyk and Summers 1992; Hanson et al. 1995; Weretilnyk et al.<br />

1995). PC may feedback-inhibit its own synthesis in sugar beet (Hanson and Rhodes<br />

1983).


6iotic Stress ~esista~ce<br />

ethanolamine @A)<br />

.<br />

e<br />

.<br />

0<br />

transferase<br />

0 SAHC<br />

0<br />

*<br />

. -0<br />

e N-<br />

H<br />

O<br />

phospho~lmonometh~lethanolamine<br />

(PIMME)<br />

phospho~ldimethylethanolamine<br />

(PDME)<br />

phospho~lcholine (PC)<br />

+-<br />

choline<br />

Figure 2 Pathway of choline synthesis in the Chenopodiaceae: The dotted line denotes feedback<br />

regulation of the first N-methyltransferase by choline.<br />

Glycinebetaine is synthesized from choline in a two-step oxidation catalyzed by a ferredoxin<br />

(Fd)-dependent choline rnonooxygenase (CMO) and a betaine aldehyde dehydrogenase<br />

(BADH; EC 1.2.1.8) with a strong preference for NAD' (Rhodes and Hanson 1993;<br />

Fig. 3).


,-,. betaine ~dehyde<br />

+ H20 - H20<br />

0 - betaine aldehyde<br />

NAD+<br />

NADH + Elt<br />

~~ycinebetai~~<br />

Pathway of glycinebetaine synthesis in the Chenopo~iacea~: Fd, fe~e~oxin.<br />

th enzymes are predominantly localized in the chloroplast stroma of chenopods<br />

et al. 1986, 1988; Brouquisse et al. 1989), although in spinach leaves approximately<br />

10% of the BADH may exist as a cytosolic isofom (~eretilnyk and Hanson<br />

1988). ~lycinebetaine is also localized predominantly in the chloroplasts of salinized<br />

spinach leaves, where it provides osmotic adjustment (Robinson and Jones 1986). Oxygen-1<br />

8 tracer studies confirm that CMO uses molecular oxygen as a substrate (Lema et<br />

al. 1988).<br />

Both CMO (Burnet et al. 1995) and BADH (~eretilnyk and Hanson 1989) have<br />

been purified to homogeneity from spinach. Complementary DNAs encoding BADH<br />

n isolated from spinach (Weretilnyk and Hanson 1990) and sugar beet (McCue<br />

son 1992). Salinity stress leads to a two- to threefold increase of CMO and<br />

ene expression, and concomitant increases<br />

enzyme level (Hanson et al. 1995).<br />

ADH has an unusual chloroplast-transit target sequence (Rathinasabapathi<br />

ansgenic tobacco plants overexpressing chloroplast-localized BADH have<br />

been obtained (Rathinasabapathi et al. 1994). However, these are unable to accumulate<br />

glycinebetaine in the absence of exogenously supplied betaine aldehyde because tobacco<br />

lacks C ~ (~athinasabapathi<br />

O<br />

et al. 1994). A cDNA encoding CMO has recently been<br />

cloned from spinach (Rathinasabapathi et al. 1997). The cDNA sequence confirms that<br />

is an Fe-S protein, and that CMO has a typical chloroplast-transit peptide sequence<br />

inasabapathi et al, 1997). Expression of chloropIast-localized CMO in BADH-con-


taining transgenic tobacco plants will provide a critical test of the adaptive value of<br />

glycinebetaine in salinity and drought stress resistance.<br />

It should be cautioned, however, that BADH may be a general aldehyde dehydrogenase,<br />

acting on other aldehyde substrates in addition to betaine aldehyde (Trossat et al.<br />

1997).BADHwilluse 3-dimethylsulfoniopropionaldehyde, an intermediate in DMSP<br />

synthesis (see Sec. 1V.B) and certain aldehydes (3-aminopropionaldehyde and 4-aminobutyraldehyde)<br />

involved in polya~ne metabolism (Trossat et al, 1997). Transgenic<br />

plants engineered for BADH expression (Holmstrom et al. 1994; ~athinasabapathi et al.<br />

1994) may possibly exhibit phenotypes related to perturbed polyamine metabolism, in<br />

addition to phenotypes attributable to alterations in the glycinebetaine synthesis pathway<br />

(Trossat et al. 1997). The multiple substrate specificities of BADH may explain the<br />

occurrence of BADH in plants that do not accumulate glycinebetaine (Ishitani et al.<br />

1993; Weretilnyk et al. 1989), and in organs of glycinebetaine-accumulating plants that<br />

do not contain glycinebetaine (e.g., roots of cereals; Ishitani et al. 1995).<br />

In monocotylednous plants, recent studies suggest that BADH is localized in peroxisomes<br />

(Nakamura et al. 1997). All monocotyledenous BADHs have a COOH-terminal<br />

tripeptide (SKI.,) that is a signal for targeting preproteins to microbodies (Nakamura et<br />

al. 1997). This raises the question of whether the subcellular localization of the choline<br />

oxidation pathway is the same in grasses as in chenopods.<br />

In bacteria, such as E. coli, choline is an osmoprotectant only because it is oxidized<br />

to glycinebetaine. Mutants in which choline fails to serve as an osmoprotectant prove to<br />

have defects in choline transport (betT), choline oxidation (betA), betaine aldehyde oxidation<br />

(betB), or a regulatory locus (betl) encoding a repressor protein (Lamark et al, 1991,<br />

1996). The entire E. coli bet gene cluster has been introduced into the freshwater cyanobacterium<br />

~y~ec~ococc~s and has conferred glycinebetaine accumulation and increased<br />

salt tolerance, partly owing to stabilization of photosystem I1 (PS-11; Nomura et al.<br />

1995a). However, this increased salt tolerance appears to depend on the presence of an<br />

exogenous supply of choline. Transgenic plants expressing the E. coli betA gene encoding<br />

choline dehydrogenase putatively capable of oxidizing both choline and betaine aldehyde,arereportedtohaveincreasedsalttolerance<br />

(Liliusetal.1996).However,no<br />

evidence was presented that these transgenic plants actually accumulated glycinebetaine.<br />

In maize, several naturally occurring glycinebetaine-de~cient inbred lines have<br />

been identified (Brunk et al. 1989). These lack the ability to accumulate glycinebetaine<br />

in leaf tissue in response to either salinity stress or water deficits owing to a single<br />

recessive gene (betl), which has been mapped to the short arm of chromosome 3 near<br />

the centromere (Rhodes et al. 1993; rang et al. 1995). Glycinebetaine deficiency in<br />

maize is associated with an inability to oxidize ['4C]choline to ~'4C]glycinebetaine, but<br />

an unimpaired ability to oxidize [2H3]betaine aldehyde to [2H3]glycinebetaine, suggesting<br />

a lesion at the CMO step in the biosynthetic pathway (Lema et al. 1991). Homozygous<br />

glycinebetaine-de~cient maize lines appear to be more salt-sensitive than near-isogenic<br />

homozygous glycinebetaine-containing lines (Saneoka et al. 1995), and they exhibit<br />

greater membrane injury and damage to PS-I1 in response to heat stress (Yang et al.<br />

1996). ~lycinebetaine-deficient maize lines exhibit a signi~cantly elevated pool of free<br />

choline; however, choline does not accumulate to levels equal to the glycinebetaine level<br />

of glycinebetaine-containing lines, suggesting that choline must down-regulate its own<br />

synthesis (Yang et al. 1995). Consistent with this, glycinebeta~ne-deficient lines of maize<br />

exhibit an elevated level of serine in comparison with homozygous glycinebetaine-containing<br />

lines (Yang et al. 1995).


Clycinebetaine is thought to be the archetypal betaine in the plant kingdom (Weretilnyk<br />

et al. 1989; Hanson and Burnet 1994; Hanson et al. 1994a). Some species have apparently<br />

lost the capacity to accumulate large quantities of this compound, and others have<br />

evolved the capacity to synthesize alternative quaternary a~onium compounds, including<br />

choline- sulfate, p-alaninebetaine, or prolinebetaine. This is particularly well<br />

illustrated in a consideration of the quaternary ammonium compounds accumulated by<br />

membersofthePlumbaginaceae (Hansonetal.1994a).Ancestralsubfamilies of the<br />

Plumbaginaceae are glycinebetaine-accumulators (Hanson et al. 1994a). In more advanced<br />

subfamilies, however, glycinebetaine has been replaced in part or in whole by p-<br />

alaninebetaine or prolinebet~ne (Hanson et al. 1994a). In addition, all members of the<br />

Plumba~inaceae accumulate choline-0-sulfate. This compound is synthesized from choline<br />

by a salt stress-inducible cholinesulfotransferase (EC 2.8.2.6; Rivoal and Hanson<br />

1994b; Fig. 4).<br />

The adaptive significance of choline-0-sulfate in the Plumbaginac~ae is thought to<br />

reside in the fact that salt glands of these species can excrete chloride, but not sulfate;<br />

thus, conjugation of sulfate with choline is proposed to be a mechanism of sulfate detoxification,<br />

converting a normally deleterious anion to a compatible osmolyte (Hanson et<br />

al. 1994a). Because choline- sulfate synthesis competes with glycinebetaine synthesis<br />

for available choline, this may have cont~buted to the evolution of alternative betaine<br />

biosynthetic pathways (p-alaninebetaine or prolinebetaine) that draw on substrates other<br />

than choline (Hanson et al. 1994a).<br />

p-Alaninebetaine is confined to species of the Plurnbaginaceae (Hanson et ai. 1994a). It<br />

is synthesized by direct ~-methylation of the amino acid p-alanine, presumably catalyzed<br />

by SAM-dependent ~-methyltransferases (Hanson et al. 1991, 1994a; Fig. 5). In contrast<br />

with glycinebetaine synthesis, p-alaninebetaine does not require oxygen (Hanson et al.<br />

1994a). It is possible, therefore, that this pathway represents an adaptation to anoxic<br />

saline environments (Hanson et al. 1994a). The ~-alaninebetaine biosynthetic pathway is<br />

HO<br />

I<br />

3-phosp~oadenosi~~-<br />

5'phosphosulfate (PAPS)<br />

3-phosphoadenosine-<br />

0 5 'phosphate (PAP)<br />

- G~oli~e-0-sulfate<br />

Pathway of choline-0-sulfate synthesis in the Plumbaginaceae.


0<br />

ll<br />

HO NW, @-alanine<br />

SAM<br />

SAHG<br />

0<br />

"0 - @-~~inebetaine<br />

ure 5 Putative pathway of P-alaninebetaine synthesis in the Plumbaginaceae.<br />

active in both leaves and roots of p-alaninebetaine-accumulating Li~oniu~ species<br />

(Hanson et al. 1991 ). The precise origin of p-alanine remains obscure. This amino acid<br />

apparently is not derived from decarboxylation of aspartate in Li~o~2i~~<br />

species (Hanson<br />

and Burnet 1994).<br />

In some higher-plant species (notably members of the Labiateae, Asteraceae, Fabaceae,<br />

Bataceae, Rutaceae, Plumbaginaceae, and Capparaceae), proline can be metabolized to<br />

prolinebetaine (stachydrine; Naidu et al. 1987, 1992; Wood et al. 1991; Hanson et al.<br />

1994a; Bonham et al. 1995; Gorham 1995; Blunden et ai. 1996; McLean et al. 1996;<br />

Nolte et al. 1997). The synthesis of prolinebetaine is thought to involve N-methylation of<br />

proline by SAM[-dependent methyltransferases through the intermediate N-methylproline<br />

(hygric acid; Robertson and Marion, 1960; Essery et al. 1962), but the enzyme(s) have<br />

not yet been characterized (Fig. 6). Hydroxyprolinebetaine often (but not always) accumulates<br />

with prolinebetaine in these prolinebetaine-accu~ulating species (Wood et al.<br />

1991; Hanson et al. 1994a; Noltet al. 1997). The proline and N-methylproline hydroxylase(s)<br />

putatively involved in hydroxyproIinebetaine synthesis have not yet been characterized.<br />

Comparisons of related species with different capacities for accumulation of prolinebetaine<br />

and hydroxyprolinebetaine suggest that the free-proline pool size is inversely<br />

related to the total prolinebetaine plus hydroxyprolinebetaine pool size (Hanson et al.<br />

1994a; Nolte et al. 1997). Prolinebetaine is a more potent osmoprotectant than proline<br />

(Hanson et al. 1994a). Conversion of proline to prolinebetaine, therefore, may enhance<br />

osmotic stress resistance. In principle, increased flux of proline to N-methylated derivatives<br />

should alleviate P5CS from feedback regulation by proline by decreasing the pool


33 o~es et a/.<br />

HO<br />

. P5C P5C<br />

.<br />

.<br />

0<br />

sA.Hc<br />

SAM<br />

SAHC<br />

6 Putative pathway of prolinebetaine biosynthesis.<br />

size of free proline (See Sec. 11; and also Chap. 8), simultaneously restricting proline<br />

oxidation and, hence, osmolyte catabolism (Samaras et al. 1995). One isomer of hydroxyprolinebetaine<br />

(~~~~~-4-hydroxy-~-prolinebetaine) is a potent inhibitor of animal acetylcholinesterase<br />

(Friess et al. 1957), raising the possibility that the ac~umulation of this<br />

compound may serve a dual function: as a compatible osmolyte and an antifeedant,<br />

affording protection against herbivores (Hanson and Burnet 1994). In contrast, proline is<br />

a stimulant of insect herbivory (Haglund 1980).<br />

Pipecolatebetaine is accumulated together with prolinebetaine in certain species; notably<br />

~e~icago and Ac~i~Zea (Wood et al. 1991; Bonham et al. 1995). Pipecolic acid is often<br />

accumulated to high levels by members of the Fabaceae (Stewart and Larher 1980; Rosenthal<br />

1982), and indeed, free pipecolic acid is a major constituent of the free amino<br />

acid pool of ~e~icugo (Wood et al. 1991). The putative pathway of pipecolatebetaine


SAM<br />

SAHC<br />

SAM<br />

SAHC<br />

ure 7 Putative pathway of pip~colatebetaine biosynthesis.<br />

synthesis is illustrated in Figure 7. It is unknown if the proline and N-methylproline N-<br />

methyltransferases have dual substrate specificity, and if they are also capable of N-<br />

methylating pipecolic acid and N-methylpipecol~c acid, respectively. Pipecolic acid is<br />

derived from lysine catabolism by ~-aminoadipic semialdehyde (Gon~alves-Butruille et<br />

al. 1996), which spontaneously cyclizes to ~'-pipe~dine-6-c~boxylate. The latter can<br />

then be reduced to pipecolic acid in a reaction analogous with that catalyzed by P5CR<br />

(Stewart and Larher 1980; Rosenthal 1982; Galili 1995).<br />

tiv<br />

The synthesis of the foregoing ~uatern~y a~monium compounds is dependent on the<br />

activity of SAM-dependent ~-methyltransferases for N-methylation of ethanolamine<br />

bases, p-alanine, proline, and pipecolic acid. This implies that osmolyte accumulation is<br />

critically dependent on the supply of SAM as a methyl donor (Hanson et al. 1995;<br />

Bohnert and Jenson 1996). When SAM is used as a methyl donor in ~-~ethyl~ansferase<br />

reaction, the resulting ~-adenosylhomocysteine (SAHC) is recycled to homocysteine<br />

(HC) by adenosylhomoc~steinase (EC 3.3.1.1). HC can then accept a methyl group from<br />

the ~5-~ethy1-tetrahy~ofolate (~~-m~thyl-THF) pool to regenerate methionine by the<br />

action of methionine synthase (EC 2.1.1.14) and, thereby, SAM by the action of S<br />

synthetase (methionine adenosyltransferase; EC 2.5.1.6) (Giovanelli et al. 1980; Fig. 8).<br />

The foregoing cycle is known as the act~v~te~ ~et~yl cycle (Bohnert and Jenson


T€€F<br />

ethionine<br />

TEE<br />

~o~ocysteine<br />

Ns-me~yl-THF (CH,-THF)<br />

FAD<br />

FADH,<br />

glycine NH, + CO, glycine serine<br />

N5,Nl0-~ethylene-T~F<br />

..<br />

N5,~l0-methe~yl-~HF<br />

formate<br />

+ THE:<br />

Alternative pathways of synthesis of ~5-methyl-tetr~ydrofolate as methyl donor for<br />

methionine synthesis from homocysteine: GDC, glycine decarboxylase; SHMT, serine hydroxymethyltransferase.<br />

hydrolase (EC 3.5.4.9), and ~5,~10-methenyl-THF dehydrogenase (EC 1.5.1 S ) occur on<br />

a single polypeptide called C1-THF synthase (Prabhu et al. 1996). In plants, however,<br />

the synthetase activity occurs on a protein moiety separate from that of the cyclohydrolase<br />

and reductase, which occur on a bifunctional protein (Prabhu et al. 1996). Because<br />

the latter reactions are reversible, this allows an equilibrium between the pools of Cl<br />

units at the formyl and methylene levels of oxidation (Cossins 1987). Therefore, serine,<br />

glycine, and formate can serve as the precursors for methionine synthesis (Cossins 1987;<br />

see Fig. 9). [13CG]Formate is metabolized to p-['3C]serine in Ar~bi~opsis in a manner<br />

consistent with C1-THF synthase and SHMT activities interacting with a common THF<br />

pool (Prabhu et al. 1996). Formate is a very effective precursor of the methyl groups of


choline and glycinebetaine in glycinebetaine-accumulating plant species (Delwiche and<br />

Bregoff 1958; Hitz et al. 1981). Formate can be derived from nonenzymatic Hz02-mediated<br />

decarboxylation of glyoxylate, although Kleczkowski and Givan (1988) question<br />

whether this pathway is of physiological significance.<br />

The sulfonium analogue of P-alaninebetaine, 3-dimethylsulfoniopropionate (DMSP), is<br />

synthesized and accumulated to osmotically significant levels in a few higher-p~ant species<br />

from diverse families; certain species of the Asteraceae (most notably WoZZasto~ia<br />

b$ora), and Poaceae (some [but not all] species of S~arti~ and Sacc~ar~~)<br />

(Paquet et<br />

al. 1994; J.ames et al. 1995b), and by many marine algae (see, e.g., Blunden and Gordon<br />

1986; Dickson and Kirst 1986; Gage et al. 1997). DMSP is as effective as glycinebetaine<br />

as an osmoprotectant for E. coli (Paquet et al. 1994), and is noted to be a potent cryoprotectant<br />

(~ishiguchi and Somero 1992; Karsten et al. 1996).<br />

In ~oZZaston~ff, DMSP is synthesized by methylation of methionine to S-methylmethionine<br />

(SMM) in the cytosol (Hanson et al. 1994b; James et al, 1995a,b), transport of<br />

SMM into the chloroplast, dec~boxylation and deamination of SMM to DMSP-aldehyde<br />

(James et al. 1995b), and oxidation of the aldehyde to DMSP in the chloroplast (Trossat<br />

et al. 1996; Fig. 30). The first enzyme of the pathway, SA~~L-methionine methylt transferase<br />

(EC 2.1.1.12) has been purified from WoZlastoni~ leaves (James et al. 1995a).<br />

The final step in the pathway is catalyzed by a chloroplast-localized DMSP-aldehyde<br />

dehydrogenase, which may be identical with the BADH involved in glycinebetaine synthesis<br />

(James et al. 1995b; Trossat et al. 1996, 1997; Vojtechovd et al. 1997). The enzymes<br />

responsible for metabolism of SMM to DMSP-aldehyde are not yet known.<br />

In marine algae, DMSP synthesis occurs by a route that is entirely different from<br />

that in higher plants (Gage et al. 1997; Fig. 11). Methionine is first transaminated to 4-<br />

methylthio-2-oxobutyrate (MTOB), which is then reduced to 4-methylthio-~-hy~oxybutyrate<br />

(MTHB). S-Methylation of MTHB then yields 4-dimethylsulfonio-2-hydroxybutyrate<br />

(DMSWB), which is subs~uently oxidatively decarboxylated to DMSP (Gage et al.<br />

1997; see Fig. 11).<br />

Because DMSP does not contain nitrogen, the accumulation of DMSP could offer<br />

advantages over accumulation of nitrogenous onium compounds for osmotic adjustment<br />

in N-li~ting environments; N-deficiency increases DMSP levels ~oZZasto~i~<br />

in (Wanson<br />

et al. 1994b) and marine algae (Gage et al. 1997). The enzyme DMSP lyase generates<br />

the gas dimethyl sulfide (DMS) and the potentially toxic compound, acrylate, from<br />

DMSP. Acrylate generated from DMSP may play a role in defense against herbivores<br />

(Wolfe et al. 1997).<br />

S-Adenosylmethionine serves as a precursor of the plant hormone ethylene, implicated<br />

in the control of numerous developmental processes (Lieberman 1979; Yang and Hoffman<br />

1984; Yang 1989; Kende 1993). SAM is converted to the amino acid 1 -aminocyclopropane-1-carboxylic<br />

acid (ACC) by ACC synthase (EC 4.4.1.14), which is then oxidized<br />

to ethylene by ACC oxidase (Kende 1993; Fig. 12).<br />

The HCN generated from the catalytic action of ACC oxidase is detoxified by rapid


‘S<br />

~~~:L-methionine<br />

~-methyltransf~ras~<br />

OH c ethionine<br />

SAM:<br />

SAHC<br />

tr~s~inatio~<br />

& dec~boxylation<br />

3-dime~ylsulfonio-<br />

(~~SP-alde~yde)<br />

Figure 10 Pathway of synthesis of dimethylsulfoniopropionate (DMSP) in ~ollffstonjff ~~orff:<br />

SAM, ~-adenosyl~ethionine; SAHC, S-adenosylhomocysteine.<br />

metabolism to the amino acid P-cyanoalanine that can then be hydrated to asparagine or<br />

conjugated to y-glut~yl-P-cyanoalanine (Yang 1989). 5’-~ethylthioadenosine is metabolized<br />

back to methionine by 5’-methyl~hio~bose and MTOB by the methionine salvage<br />

pathway (Miyazaki and Yang 198’7; see Fig. 12). ACC synthase may be regulated by its<br />

substrate SAM through S~-induced inactivation. This mechanism may be responsible<br />

for the short half-life of this enzyme in vivo (Yang 1989). ACC oxidase requires Fe2”<br />

for catalytic activity and, in some systems, is activated by C02 (Kende 1993). The plant<br />

species ~ ot~~ogeto~ ~ecti~at~s is unusual in that it lacks ACC oxidase and ethylene<br />

biosynthesis; this species is consequently rich in ACC (Summers et al. 1996).<br />

1 -<strong>Amino</strong>cyclopropane- 1 -carboxylate synthase is induced by various environmental<br />

stresses, including mechanical stress (wounding or mechanical impedance), drought,<br />

chilling, anoxia, and hypoxia (Yang 1989; Kende 1993; He et al. 1996a). ACC oxidase<br />

isalsoinduced by hypoxia and by mechanical impedance in maize roots (He et al.<br />

1996a); however, with the combination of hypoxia plus mechanical impedance, ethylene<br />

formation and ACC synthase induction exhibit a greater synergism than ACC oxidase<br />

induction, cellulase induction, and aerenchyma formation (He et al. 1996a). ~erenchyma<br />

formation accelerates the transfer of O2 from the aerial tissues to the ~2-de~cient tissues


ho~es et al.<br />

4-methylthio-2-<br />

OH hydro~bu~rate (MTHB)<br />

OH<br />

4-dime~ylsulfonio-2-<br />

hydro~ybu~rate ( ~ ~ S ~ )<br />

ure 11 Proposed pathway of synthesis of di~ethylsulfoniopropionate (DMSP) in marine<br />

algae.<br />

of the stem base and root of flooded plants (He et al. 3996b). Ethylene production in<br />

response to hypoxia is implicated in the signal transduction pathway leading to cell death<br />

and lysis in the root cortex resulting in aerenchyma formation in maize roots (He et al.<br />

1996b; Drew 1997). Because ACC oxidase is oxygen-dependent, ACC can accumulate<br />

markedly in roots exposed to anoxia. In flooded tomato plants, the accumulated ACC is<br />

transported from the root to the shoot (through the xylem) where it is then rapidly oxidized<br />

to ethylene, inducing epinasty (Bradford and Yang 1980; English et al, 1995) and<br />

adventitious root formation (Visser et al. 1996). The latter may partly involve an ethylene-mediated<br />

increase in the sensitivity of root-forming tissue to endogenous indoleacetic<br />

acid (Visser et al. 1996).<br />

ACC can be conjugated to malonyl-ACC (MACC). MACC synthesis is catalyzed<br />

by ACC ~-~alonyltransferase (Yang1989;MartinandSaftner1995). An alternative<br />

ACC conjugate has recently been identified in tomato fruits: (y-gluta~yla~ino)-cyclo-


337<br />

ATP<br />

I<br />

\<br />

methio~ine<br />

salvage<br />

pathway<br />

PPi + Pi<br />

'S<br />

I<br />

Ad<br />

Smethyl-<br />

~hioadenosin~<br />

(rnA.1<br />

OH<br />

C,Hd + C02 + HCN + 2H20<br />

Figure 12 Pathway of ethylene (C21-&) biosynthesis in plants, and pathway of recycling of 5'-<br />

methylthioadenosine (MTA) to methionine (the methionine salvage cycle) in plants: ACC, 1-<br />

aminocyclopropane-1-carboxylate; MTR, ~ethylthio~bose; MTOB, 4-methylthio-2-oxobutyrate;<br />

SAMS, ~-adenosylmethionine synthetase.<br />

propane-1-carboxylic acid (GACC; Martin et ai. 1995). MACC and GACC formation<br />

may contribute to the regulation of ACC levels and, thereby, of ethylene formation (Yang<br />

1989; Kende 1993; Banga et al. 1996).<br />

The diamine putrescine, the trimine spermidine, and the tetramine spermine are ubiquitous<br />

in plant cells (Smith et al. 1979; Bagni and Pistocchi 1992). They occur as free<br />

cations and as conjugates with phenolic acids and macromolecules (Galston and Sawnhey<br />

1990). Their levels increase greatly in response to environmental stresses, most notably<br />

under conditions of potassium deficiency, water deficits, salinity stress, anaerobiosis,<br />

and acid stress (Flores et al. 1989). Because polyamines are synthesized by amino acid<br />

decarboxylation reactions that consume €I?, polyamine accumulation may function as<br />

part of a homeostatic mechanism to keep intracellular pH at a constant value (Flores et<br />

al. 1985). Polyamines may also play a role in the regulation of DNA replication and cell<br />

division, and they are implicated in the control of senescence and morphogenesis (Evans<br />

and Malrnberg 1989; Galston and Sawnhey 1990).<br />

Arginine decarboxylase (ADC; EC 4.1.1.19)-(a chloroplast-localized enzyme;<br />

Borrell et al. 1995)"is induced by a variety of stresses (most notably potassium deficiency;<br />

Watson and Malmberg 1996) and is thought to be the enzyme primarily responsible<br />

for environmental stress-induced putrescine accumulation (Galston and Sawnhey


199~; Fig. 13). ADC of oats is clipped from a precursor into two polypeptides found in<br />

the soluble enzyme (Malmberg et al. 1992). The two polypeptides are linked by disul~de<br />

bonds to form the active enzyme (Watson and Malmberg 1996). The oat (Bell and Malmberg<br />

1990), tomato (Rastogi et al. 1993), and Ar~bi~o~si~ atso son and Malmberg 1996).<br />

C genes have been cloned. In Ar~bi~o~sis, ADC enzyme levels do not correspond<br />

to ADC protein amounts, suggesting post~anslational modification of the enzyme<br />

in response to potassium deficiency (Watson and Malmberg 1996). In osmotically<br />

stressed oat leaves, spermine inhibits posttranslational processing of the ADC precursor,<br />

with subsequent decreases in mature ADC (Borrell et al. 1996).<br />

a~~atine<br />

~athways of synthesis and metabolism of putrescine.


The alternative, more direct, pathway of synthesis of putresine through ornithine<br />

decarboxylation catalyzed by cytosolic ornithine decarboxylase (ODC; EC 4" 1.1.17; see<br />

Fig. 13) is proposed to be of little importance in stress-induced putrescine accumulation,<br />

but may be critical in regulation of developmental processes (Galston and Sawhney 1990;<br />

Walden et al. 1997). Increased putrescine biosynthesis, mediated by transgenic ODC,<br />

promotes somatic embryogenesis in carrots (Bastola and Minocha 1995). The diamine<br />

cadaverine, derived from the amino acid lysine by decarboxylation, may play an important<br />

role in root development (Gamarnik and Frydrnan 1991).<br />

The triamine spermidine is synthesized by condensation of putrescine with decarboxylated<br />

SAM, synthesized by SAM decarboxylase (EC 4.1.1.50; Flores et al. 1989):<br />

SAM + Ht SAM decarboxylase<br />

> decarboxylated SAM + C02<br />

The condensation of decarboxylated SAM and putrescine is catalyzed by spermine<br />

synthase (putrescine aminopropyltransferase; EC 2.5.1.16). Further condensation of spermidine<br />

with decarboxylated SAM, catalyzed by spermine synthase (EC 2.5.1.22), produces<br />

the tetramine, spermine (Flores et al. 1989; see Fig. 13). Recently, the assumption<br />

that spermidine synthase and spermine synthase are two separate enzymes in plants, has<br />

been questioned (Bagga et al. 1997). The putrescine aminopropyltransferase of alfalfa<br />

may be responsible not only for spermidine and spermine synthesis, but also the synthesis<br />

of the uncommon polyamines, thermospermine, homocaldopentamine, and homocaldohexamine<br />

(Bagga et al. 1997).<br />

Antisense SAM decarboxylase potato plants show markedly altered phenotypes,<br />

including stunted growth; short internodes; stem branching and small leaves; decreased<br />

SAM decarboxylase transcripts; a reduction in SAM decarboxylase activity; reductions<br />

in the levels of putrescine, Spermidine, and spermine; and a marked increase in ethylene<br />

production (Kumar et al. 1996). This supports the idea that there is a competitive interaction<br />

between polyamine and ethylene biosynthesis (Walden et al. 1997).<br />

In addition to serving as a precursor of spermidine and spermine, putrescine has<br />

three other metabolic fates: conjugation with hydroxycinnamic acids, metaboli~m to y-<br />

aminobutyrate (GABA), and utilization in alkaloid biosynthesis, as indicated in Figure<br />

13. Hydroxycinnamic amide (HCA) conjugates of polyamines accumulate markedly in<br />

the floral apex during flower development and are implicated in the development of<br />

competence to flower; certain mutants of tobacco that are deficient in HCAs are unable<br />

to flower, and male sterile mutants of maize lack HCA accumulation in anthers (Flores<br />

et al. 1989). A novel polyamine conjugate, ~-hexanoylspermidine, has recently been<br />

identified in senescing pea ovaries and petals (Perez-Amador et al. 1996). GABA can be<br />

derived from putrescine (through y-aminobutyraldehyde) through the reactions catalyzed<br />

by diamine oxidase and y-aminobutyraldehyde dehydrogenase (Flores et al, 1989). As<br />

noted in Section III.B, the latter enzyme may be the same as BADH involved in glycinebetaine<br />

synthesis (Trossat et al. 1997). Putrescine serves as precursor of the nicotine and<br />

tropane alkaloids (Smith et al. 1979; Flores et al. 1989; Hashimoto and Yamada 1994),<br />

which may play important roles in plant defense against herbivores (see Fig. 13).<br />

The tripeptide glutathione (GSH), is synthesized from the amino acids glutamate, cysteine,<br />

and glycine by the enzymes y-glut~ylcysteine synthetase (EC 6.3.2.2) and glutathione<br />

synthetase (EC 6.3.2.3; Meister 1983):


Glutamate + cysteine + ATP<br />

~Glutamylcysteine<br />

synthetase<br />

> y-glutamylcysteine + ADP + Pi<br />

y-glutamylcysteine + glycine + ATP<br />

synthetase > glutathione + ADP + Pi<br />

Glutathione (GSH)playsa key role in detoxification of certainxenobioticsin<br />

higher plants through the action of glutathione S-transferases (EC 2.5.1 I) 18; Breaux 1986;<br />

Pickett and Lu 1989; Rossini et al. 1996). The glutathione conjugate of a xenobiotic can<br />

then be transported into and sequestered by the vacuole (Li et al. 1995; Kreuz et al.<br />

1996; Marrs 1996). Glutathione S-transferases are induced by a wide range of chemical<br />

agents (Ulmasov et al. 1995), wounding, heavy metals, ethylene, and ozone (Marrs<br />

1996). These enzymes may have initially evolved to conjugate naturally occurring compounds,<br />

such as phenylpropanoids (Dean et al. 1995), and naturally occurring endogenous<br />

reactive electrophilic compounds (Marrs 1996), but may have then been recruited<br />

to detoxify anthropogenic exogenous xenobiotics (Kreuz et al, 1996; Mans 1996).<br />

Under a variety of environmental stresses (e.g., high light and low temperature,<br />

water deficits, or others that disrupt membrane-bound electron transport systems), plants<br />

produce a range of active oxygen species, including superoxide, hydrogen peroxide, and<br />

the hydroxyl radical (Smirnoff 1993). Superoxide is converted to hydrogen peroxide by<br />

the action of superoxide dismutase. The hydroxyl radical can then form from the reaction<br />

of hydrogen peroxide with superoxide, particularly in the presence of iron or other transition<br />

state metals (Srnirnoff 1993). These toxic species can also form when plants are<br />

exposed to ozone, SO2, NO, and NO2. They react with membranes, proteins, and nucleic<br />

acids (Smirnoff 1993). Glutathione is involved in quenching these free radicals through<br />

the ascorbate-GSH cycle (Alscher 1989; Smirnoff 1993, 1995). Thus, hydrogen peroxide<br />

can be detoxified by the action of ascorbate peroxidase (EC 1.11.1.11), dehydroascorbate<br />

reductase (EC 1.8.5.1), and glutathione reductase (EC 1.6.4.2):<br />

Ascorbate + H202 Ascorbate peroxidase > H20 + dehydroascorbate<br />

~ehydroascorbate<br />

Dehy~oascorbate<br />

reductase<br />

+ GSH ascorbate + oxidized glut at hi on^ (GSSG)<br />

GSSG + NADPH + H+ Glutathione reductase<br />

> GSH + NADP'<br />

Strictly speaking, the initial product of the ascorbate peroxidase reaction is monodehydroascorbate,<br />

which can then either be reduced back to ascorbate by the action of an<br />

NAD(P)H-dependent monodehydroascorbate reductase (EC 1.6.5,4), or directly reduced<br />

back to ascorbate by an electron from PS-I (Smirnoff 1995). Any monodehydroascorbate<br />

not so reduced disproportionates to ascorbate plus dehydroascorbate (Smirnoff 1995); it<br />

is only the latter that is reduced by glutathione in the reactions depicted in the foregoing.<br />

Specific isoforms of the enzymes of this cycle are chloroplast-localized and are induced<br />

by an array of environmental factors (Alscher 1989; Pastori and Trippi 1992; Smirnoff<br />

1995). Overexpression of glutathione reductase leads to an increase in antioxidant capacity<br />

and resistance to photoinhibition in poplar (Foyer et al. 1995). However, overproduction<br />

of ascorbate peroxidase in the tobacco chloroplast does not appear to provide protection<br />

against ozone (Torsethau~en et al. 199'7).


Although the chloroplast-localized ascorbate-glutathione cycle is well characterized,<br />

it is now becoming clear that the enzymes of this cycle are also found in mitochondria<br />

and peroxisomes, and they may represent an important antioxidant protection system<br />

against H203 generated in these organelles (Jimhez et al. 1997).<br />

Glutathione transferases may also be involved in conjugating electrophiles generated<br />

from the reaction of hydroxyl radicals with lipids and DNA and, thus, contribute to<br />

oxidative stress resistance (Marrs 1996). Apoplastic GST induction may be mediated by<br />

an oxidative signal in soybean (Flury et al. 1996).<br />

F. P ~ y ~ o c ~ ~<br />

of Heavy<br />

Glutathione serves as a precursor of peptides known as' phytochelatins9 which are composed<br />

of two or more repeating y-glutamylcysteine units, with a terminal glycine residue;<br />

(~-glutamylcysteine)~-gly, where n = 2-11 (Steffens et al, 1986; Reese and ~agner<br />

1987). The enzyme responsible for the synthesis of these peptides is known as phytochelatin<br />

synthase (Rauser 1990, 1995). Phytochelatins (PCs) play an important role in the<br />

detoxification of certain heavy metals (p~ticularly cad~ium) in plants (Rauser 1990,<br />

1995; Howden et al. 1995a,b). Phytochelatin synthase is activated by cadmium (Rauser<br />

1995). In certain plants (notably legumes) that can synthesize homoglutathione, in which<br />

g-alanine is substituted for glycine as the terminal amino acid, homophytochelatins are<br />

synthesized along with PCs in response to Cd (Klapheck et al, 1995).Inmaizeand<br />

certain other species of the Poaceae, a third family of phytochelatins has been found in<br />

which serine is the COOH-te~inal amino acid (Rauser and ~euwly 1995). ~admium<br />

detoxification may require transport of the Cd-phytochelatin complexes into the vacuole;<br />

a transport system has been recently described for these complexes (Salt and Rauser<br />

1995).<br />

In Rubia tinctoru~, phytochelatins (class I11 metallothionein) are induced by many<br />

metal ions, but only a few (Ag, Cd, and Cu) are bound to the PCs that they induce<br />

(Maitani et al. 1996).<br />

v. A<br />

A<br />

During anaerobic stress, pyruvate must be metabolized to either lactate (by lactate dehydrogenase;<br />

LDH) or ethanol (by pyruvate decarboxylase [PDC] and alcohol dehydrogenase<br />

[ADH]) to regenerate the NAD" required to sustain glycolysis and associated ATP<br />

production (Davies 1980). Lactate fermentation may cont~bute to the acidification of the<br />

cytoplasm (Roberts et al, 1984; Ratcliffe 1995), which may progressively inhibit LDH<br />

and activate PDC, promoting alcohol fe~entation (Rivoal and Hanson 1994a). Accumulation<br />

of pyruvate under anaerobic conditions may shift the equilibrium of alanine aminotransferases<br />

toward alanine accumulation; a rapid response of plants to anaerobic stress<br />

(Streeter and Thompson 1972; Stewart and Larher 1980):<br />

Pyruvate + amino acid 3 alanine + 2-keto acid


An alanine a~notransferase (g1utamate:pyruvate a~notransferase; EC 2.6.1.2) is induced<br />

under anaerobic conditions in barley roots (Good and Crosby 1989; Good and<br />

uench 1992). The precise function of alanine accumulation under anoxia is unknown,<br />

butitisspeculatedthatit may serve as a storage form of pyruvate (perhaps in the<br />

vacuole), controlling supply of pyruvate to LDH and PIX and, thereby, the flux to<br />

lactate and ethanol (Good and Crosby 1989; Good and Muench 1992). The synthesis of<br />

alanine may occur at the expense of the acidic amino acids, glutamate and aspartate<br />

(Streeter and Thompson 1972; Stewart and Larher 1980), and occurs concomitantly with<br />

the accumulation of GABA (see following Section VI).<br />

.<br />

The accumulation of the amino acid y-~nobutyrate (GABA) is markedly stimulated<br />

under anaerobic conditions (Streeter and Thompson, 1972; Stewart and Larher 1980;<br />

Aurisano et al. 1995; Ratcliffe 1995; Drew 1997). Although G A can be derived from<br />

putrescine (see Sec. IV.D), the anaerobic stress-induced accumulation of GABA is<br />

thought to result primarily from the catalytic action of glutamate decarboxylase (EC<br />

4.1.1.15). In part, the stimulation of glutamate decarboxylase may be the consequence of<br />

cytoplas~c acidification; glutamate decarboxylase has an acid pH optimum (-5.8; Patterson<br />

and Graham 1987):<br />

Glutamate + H+ ‘lutarnate<br />

> GABA + CQ2<br />

ause this reaction is proton-consuming, it could represent an adaptive response conuting<br />

to regulation of cytoplasmic pH (Patterson and Graham 1987; Crawford et al,<br />

atcliffe 1995). Crawford et al. (1994) have shown that weak acids causing cytoplasmic<br />

acidification also induce GABA accumulation, and they conclude that this response<br />

is consistent with a role for GABA synthesis in active pH regulation.<br />

y-A~nobutyrate can be transaminated with pyruvate to yield alanine (or with 2-<br />

oxoglutarate to yield glutamate), generating succinic semialdehyde, which can then be<br />

metabolized to succinate by the action of succinate semialdehyde dehydrogenase (Vandewalle<br />

and Qlsson 1983; Patterson and Graham 1987; Shelp et al. 1995; Fig. 14). The<br />

latter enzymes are mitochondrial (Hear1 and Churchich 1984; Breitkreuz and Shelp 1995)<br />

and have alkaline pH optima (Patterson and Graham 1987). GABA accumulation promoted<br />

by cytosolic acidi~cation may partly result from inhibition of GABA-transaminases.<br />

The conversion of glutamate to succinate by the action of glutamate decarboxylase,<br />

GABA ~ansa~nases, and succinic semialdehyde dehydrogenase is known as the<br />

GABA shunt (Vandewalle and Olsson 1983; Patterson and Graham 1987; Breitkreuz and<br />

Shelp 1995; Shelp et al, 1995), affording an alternative pathway for glutamate entry into<br />

the TCA cycle (see Fig. 14).<br />

The accumulation of GABA is induced<br />

response to a sudden decrease in temperature<br />

(Wallace et al. 1984; Patterson and Graham 1987), heat shock (Mayer et al. 1990),<br />

mechanical manipulation (Wallace et al. 1984), and water stress (Rhodes et al. 1986).<br />

Rapid GABA accumulation in response to wounding may play a role in plant defense<br />

against insects (Ramputh and Brown 1996).<br />

Glutamate decarboxylase is a cytosol-localized enzyme (Breitkreuz and Shelp


glutamate<br />

NH,<br />

2-oxo<br />

HO<br />

succinate<br />

Figure 14 Pathway of synthesis and metabolism of GABA.<br />

1995) and has recently been shown to be a calmodulin-binding protein that is<br />

C~~al~odulin activated (Baum et al. 1993; Ling et al. 1994; Baum and Fridmann<br />

1996; Arazi et al. 1995; Snedden et al. 1995). Crawford et al. (1994) note that reduced<br />

cytosolic pH values increase Ca”* levels, and rapid and transient increases in Ca”’ levels<br />

occur in response to mechanical stress and cold stress; conditions that elicit rapid GABA<br />

accumulation (Wallace et al. 1984). The Ca”+-calmodulin activation of glutamate decarboxylase<br />

provides a link between intermedi~ amino acid metabolism and perturbations<br />

of cytosolic ea2+ (Ling et al. 1994; Baum and Fridmann 1996; Arazi et al. 1995; Snedden<br />

et al. 1995) that regulate a host of other metabolic activities (Allan and Trewavas 1985;<br />

Bush 1995).<br />

The only other enzyme of glutamate metabolism currently known to be stimulated<br />

by Ca” in plants in glutamate dehydrogenase (GDH; EC 1.4.1.2), a ~itochondrial enzyme<br />

(Turano et al. 1997):<br />

NH~ + 2-oxoglutarate + NADH + H+ < dehydrogenase > glutamate + NAD+<br />

A calcium-binding domain has been identified in the f3-subunit of GDH encoded<br />

by GDH2, but not in the a-subunit of GDH encoded by GDHI in Arabidopsis, suggesting<br />

that the different isoforms of the hexameric GDH composed of different combinations<br />

of subunits, may be differentially regulated by Ca” (Turano et al. 1997).


We have attempted to highlight areas of amino acid metabolism that may play an important<br />

role in plant “stress” resistance, by osmotic adjustment and the accumulation of<br />

compatible osmolytes; the detoxi~cation of active oxygen sp<br />

heavy metals; and intracellular pH regulation. The advances since 1980 have been primarily<br />

in the cloning of the genes encoding key enzymes of these biosynthetic pathways<br />

and the characterization of their regulation in relation to environmental stimuli. As noted<br />

by Bohnert et al. (1993, the use of metabolic engineering holds great promise for testing<br />

hypotheses conce~ing the role of radical scavenging and osmolyte accumulation as<br />

stress-protection mechanisms (e.g., Foyer et al. 1995; Kishor et al. 1995). These traits<br />

are not restricted to amino acid metabolism; significant advances have also been achieved<br />

in engineering of superoxide dismutase, and the accumulation of polyols and fructans as<br />

compatible solutes or radical scavengers that may limit injury (Bohnert et al. 1995; Shen<br />

et al. 1997, and references cited therein), These traits may eventually need to be combined<br />

with modified ion uptake, exclusion or co~partmentation, facilitated water permeability,<br />

molecular chaperones, modified membrane properties, or stress-signaling pathways<br />

to realize their full potential (Bohnert et al. 1995; Bohnert and Jensen 1996). The<br />

next decade is likely to witness further exciting progress in these areas. In the testing of<br />

transgene effects on plant resistance to low water potentials and salinity stress, we urge<br />

(1) careful measurements of growth and water relations to quantify the degree of stress<br />

and to assess potential growth penalties of transgene expression, (2) consideration of<br />

relevant fluxes in addition to pool sizes, and (3) consideration of the metabolism of<br />

growing regions as well as mature tissues.<br />

Adams, E. and Frank, L. (1980). Metabolism of proline and the hydroxyprolines. Annu. Rev. Biochern.,<br />

49: 1005-1061.<br />

Allan, E. E;. and Trewavas, A. J. (1987). The role of calcium in metabolic control. The ~joche~jstry<br />

of Plants, Vol. 12 (D. D. Davies, ed.), Academic Press, New York, pp. 117-149.<br />

Alscher, R. G. (1989). Biosynthesis and antioxidant function of glutathione in plants. P~ysjoZ<br />

Plant., 77 457-464,<br />

Anthoni, U., Christophersen, C., Hougaard, L., and Nielsen, P. H. (1991). Quate~ary ammonium<br />

compounds in the biosphere-an example of a versatile adaptive strategy. Cornp. Bioche~z.<br />

Bi~~hys~, 99B: 1-18.<br />

Arazi, T., Baum, G,, Snedden, W. A,, Shelp, B. J., and Fromm, H. (1995). Molecular and biochemical<br />

analysis of calmodulin interactions with the calmodulin-binding domain of plant glutamate<br />

decarboxylase. Plant Physiol., 108: 551-561.<br />

Aurisano, N., Bertani, A,, and Reggiani, R. (1995). Anaerobic accumulation of 4-ami~obutyrate in<br />

rice seedlings; causes and significance. P~ytoche~istry, 38: 1147-1 150.<br />

Bagga, S,, Rochford, J., Klaene, Z,, Kuehn, G, D., and Phillips, 6. C. (1997). Putrescine aminopropyltransferase<br />

is responsible for biosynthesis of spermidine, spermine, and multiple uncommon<br />

polyamines in osmotic stress-tolerant alfalfa. Plant P~ysjol., 114: 445454,.<br />

Bagni, N. and Pistocchi, R. (1992). Polyamine metabolism and compa~mentation in plant cells.<br />

~jt~ogen ~etaboljsrn of Plants (K. Mengelm and D. J. Pilbeam, eds.), Clarendon Press,<br />

Oxford, pp. 229-248,<br />

~allantyne, J. S, and Chamberlin, M. E. (1994). Regulation of cellular amino acid levels. Cellular


and Molecular Physiology of Cell Volume Regulation (K. Strange, ed.), CRC Press, Boca<br />

Raton, FL, pp. 11 1-122.<br />

Banga, M., Slaa, E. J., Blom, C. W, P. M,, and Voesenek, L. A. C. J. (1996). Ethylene biosynthesis<br />

and accumulation under drained and submerged conditions: A comparative study of two<br />

Rumex species, Plant Physiol., 112: 229-237.<br />

Barnett, N. M, and Naylor, A. W. (1966). <strong>Amino</strong> acid and protein metabolism in Bermuda grass<br />

during water stress. Plant Physiol., 41: 1222-1230.<br />

Bastola, D. R. and Minocha, S. C. (1995). Increased putrescine biosynthesis through transfer of<br />

mouse ornithine decarboxylase cDNA in carrot promotes somatic embryogenesis. Plant<br />

Physiol., 109: 63-7 1.<br />

Baum, G., Chen, Y., Arazi, T., Takatsuji, H., and Fromm, H, (1993). A plant glutamate decarboxylase<br />

containing a calmodulin binding domain. Cloning, sequence, and functional analysis.<br />

J. Biol. Chem., 268: 19610-19617.<br />

Baum, G. and Fridmann, Y. (1996). Calmodulin binding to glutamate decarboxylase is required<br />

for regulation of glutamate and GABA metabolism and normal development in plants.<br />

EMB~ J., 15: 2988-2996.<br />

Bell, E. and Malmberg, R, L. (1990). Analysis of a cDNA encoding arginine decarboxylase from<br />

oat reveals similarity to the Escheric~zia coli arginine decarboxylase and evidence of protein<br />

processing. Mol. Gen. Genet., 224: 431-436.<br />

Binzel, M. L., Hasegawa, P, M., Rhodes, D., Handa, S., Handa, A. K., and Bressan, R, A. (1987).<br />

Solute accumulation in tobacco cells adapted to NaC1. Plant Physiol., 84: 1408-1415.<br />

Blunden, G. and Gordon, S. M. (1986). Betaines and their sulphonio analogues in marine algae.<br />

Prog. Phycol, Res., 4: 39-80.<br />

Blunden, G., Yang, M.-H., Yuan, 2.-X., Smith, B, E., Patel, A,, Cegarra, J. A., Mathe, I., Jr., and<br />

Janicsak, G. (1996). Betaine distribution in the Labiatae, Biochem. Sys. Ecol., 24: 71-81.<br />

Boggess, S. F., Aspinall, D., and Paleg, L. G. (1976). Stress metabolism. IX. The significance of<br />

end-product inhibition of proline biosynthesis and of compartmentation in relation to stressinduced<br />

proline accumulation, Aust. J. Plant Physiol., 3: 513-525.<br />

Boggess, S. F. and Stewart, C. R. (1980). The relationship between water stress induced proline<br />

accumulation and inhibition of protein synthesis in tobacco leaves. Plant Sci. Letts., 17<br />

245-252.<br />

Bohnert, H. J. and Jensen, R. G. (1996). Strategies for engineering water-stress tolerance in plants.<br />

Trends Biotechnol., 14: 89-97.<br />

Bohnert, H. J., Nelson, D. E., and Jensen, R. G. (1995). Adaptations to environmental stresses.<br />

Plant Cell, 7 1099-1 111.<br />

Bonham, C. C., Wood, E(. V., Yang, W.-J., Nadolska-Orczyk, A., Samaras, Y., Gage, D, A.,<br />

Poupart, J., Burnet, M,, Hanson, A. D., and Rhodes, D. (1995). Identification of quaternary<br />

ammonium compounds byplasma desorption mass spectrometry. J. Mass Spectrorn., 30:<br />

1187-1 194.<br />

Borrell, A., Besford, R. T., Altabella, T., Masgrau, C., and Tiburcio, A. F. (1996). Regulation of<br />

arginine decarboxyIase by spermine in osmotically stressed oat leaves. Physiol. Plant., 98:<br />

105-1 10.<br />

Borrell, A., Culiafiez-Macia, F. A., Altabella, T., Besford, R. T., mores, D., and Tiburcio, A. F.<br />

(1995). Arginine decarboxylase is localized in chloroplasts. Plant Physiol., 109 771-776.<br />

Bradford, K. J. and Yang, S. F. (1980). Xylem transport of 1 -aminocyclopropane- 1-carboxylic<br />

acid, an ethylene precursor, in waterlogged tomato plants. Plant Physiol., 65: 322-326.<br />

Breaux, E. J. (1986). Identi~cation of the initial metabolites of acetochlor in corn and soybean<br />

seedlings. J. Agric. Food Chem., 34: 884-888.<br />

Breitkreuz, K. E. and Shelp, B, J. (1995). Subcellular compartmentation of the 4-aminobutyrate<br />

shunt in protoplasts from developing soybean cotyledons. Plant Physiol., 108: 99-103.<br />

Briens, M. and Larher, F. (1982). Osmoregulation in halophytic higher plants: A comparative study<br />

of soluble carbohydrates, polyols, betaines and free proline. Plant Cell Environ., 5: 287-292.<br />

Brouquisse, R,, Weigel, P., Rhodes, D., Yocum, C. F,, and Hanson, A. D. (1989). Evidence for a


ferredoxin-dependent choline monooxygenase from spinach chloroplast stroma. Plant Physiol.,<br />

90: 322-329.<br />

Brunk, D. G., Rich, P. J., and Rhodes, D. (1989). Genotypic variation for glycinebetaine among<br />

public inbreds of maize. Plant Physiol., 91: 1 122-1 125.<br />

Burnet, M., Lafontaine, P. J., and Hanson, A. D. (1995). Assay, purification, and partial characterization<br />

of choline monooxygenase from spinach. Plant Physiol., 108: 581-588.<br />

Bush, D. S. (1995). Calcium regulation in plant cells and its role in signalling. Annu. Rev. Plant.<br />

Physiol. Plant Mol. Biol., 46: 95-122.<br />

Cossins, E. A. (1987). Folate biochemistry and the metabolism of one-carbon units. The Biochemistry<br />

of Plants, Vol. 11 (D. D. Davies, ed.), Academic Press, New York, pp. 317-353.<br />

Crawford, L. A., Bown, A. W., Breitkreuz, K. E., and Guinel, F. C. (1994). The synthesis of y-<br />

aminobutyric acid in response to treatments reducing cytosolic pH. Plant Physiol., 104:<br />

865-87 1.<br />

Csonka, L. N. and Baich, A. (1983). Proline biosynthesis. <strong>Amino</strong> <strong>Acid</strong>s, Bios~nt~zesis and Genetic<br />

Regulation (K. M. Herrmann and R. L. Somerville, eds.), Addison-Wesley, Reading, MA,<br />

pp. 35-51.<br />

Datko, A. H. and Mudd, S, H. (1988a). Phosphatidylcholine synthesis: Differing patterns in soybean<br />

and carrot. Plant Physiol., 88: 854-861.<br />

Datko, A. H. and Mudd, S. H. (1988b). Enzymes of phosphatidylcholine synthesis in Lemna,<br />

soybean and carrot. Plant Physiol., 88: 1338-1348.<br />

Davies, D. D. (1980). Anaerobic metabolism and the production of organic acids. The Biochemistry<br />

of Plants, Vol. 2 (D. D. Davies, ed.), Academic Press, New York, pp. 581-61 1.<br />

Dean, J. V., Devarenne, T, P., Lee, I.-S., and Orlofsky, L. E. (1995). Properties of a maize glutathione<br />

S-transferase that conjugates coumaric acid and other phenylpropa~oids. Plant Physiol.,<br />

108: 985-994.<br />

Delauney, A. J. and Verma, D. P. S. (1990). A soybean ~'-pyrroline-5-carboxylate reductase gene<br />

was isolated by functional complementation in Escherichia coli and is found to be osmoregulated.<br />

Mol. Gen. Genet., 221: 299-305.<br />

Delauney, A. 3. and Verma, D, P. S. (1993). Proline biosynthesis and osmoregulation in plants.<br />

Plant J., 4: 215-223.<br />

Delwiche, C. C. and Bregoff, H. M. (1958). Pathway of betaine and choline synthesis in Beta<br />

vulgaris. J. Biol. Chem., 233: 430-433.<br />

Dickson, D. M. J. and Kirst, G. 0. (1986). The role of ~-dimethylsulphoniopropionate, glycine<br />

betaine and homarine in the osmoacclimation of Pla~monas su~cord~o~~zis. Planta, 167<br />

536-543.<br />

Drew, M. C. (1997). Oxygen deficiency and root metabolism: Injury and acclimation under hypoxia<br />

and anoxia. Annu. Rev. Plant Physiol. Plant Mol. Biol,, 48: 223-250.<br />

Elthon, T. E. and Stewart, C. R. (1982). Proline oxidation in corn mitochondria. Plant Physiol.,<br />

70: 567-572.<br />

English, P. J., Lycett, 6. W., Roberts, J. A, and Jackson, M. B. (1995). Increased l-aminocyclopropane-1-carboxylic<br />

acid oxidase activity in shoots of flooded tomato plants raises ethylene<br />

production to physiologically active levels. Plant Physiol., 109: 1435-1440.<br />

Espartero, J., Pintor-Toro, J. A., and Pardo, J. M. (1994). Differential accumulation of S-adenosylmethionine<br />

synthetase transcripts in response to salt stress. Plant Mol. Biol., 25: 217-227.<br />

Essery, J. M., McCaldin, D. J., and Marion, L. (1962). The biogenesis of stachydrine, Ph~tochemistry,<br />

1: 209-213,<br />

Evans, P. T. and Malmberg, R. L. (1989). Do polyamines have roles in plant development? Annu.<br />

Rev. Plant Physiol. Plant Mol. Biol., 40: 235-269.<br />

Flores, H. E., Protacio, C. M., and Signs, M. W. (1989). Primary and secondary metabolism of<br />

polyamines in plants. Plant Nitrogen Metabolis~ (J. E. Poulton, J. T. Romeo, E. E. Conn,<br />

eds.), Recent Adv. Phytochem., 23: 329-393.<br />

Flores, H. E,, Young, N. D., and Galston, A. W. (1985). Polyamine metabolis~ and plant stress.


347<br />

Cellular and ~olecular Biology of Plant Stress (J. L. Key and T. Kosuge, eds.), Alan R.<br />

Liss, New York, pp. 93-1 14.<br />

Flury, T., Wagner, E., and Kreuz, K. (1996). An inducible glutathione S-transferase in soybean<br />

hypocotyl is localized in the apoplast. Plant Physiol., 112: 1185-1 190.<br />

Forlani, G., Scainelli, D., and Nielsen, E. (1997). h'-Pyrroline-5-carboxylate dehydrogenase from<br />

cultured cells of potato. Plant Physiol., 113: 1413-1418.<br />

Foyer, C. H., Souriau, N., Perret, S., Lelandais, M., Kunert, K.-J., Pruvost, C., and Jouanin, L.<br />

(1995). Overexpression of glutathione reductase but not glutathione synthetase leads to increases<br />

in antioxidant capacity and resistance to photoinhibition in poplar trees. Plant Physiol.,109:<br />

1047-1057.<br />

Friess, S. L., Patchett, A. A., and Witkop, B. (1957). The acetylcholineesterase surface. VII. Interference<br />

with surface binding as reflected by enzymatic response to turicine, betonicine and<br />

related heterocycles. J. Am. Chem. Soc., 79: 459-462.<br />

Fukutoku, Y. and Yamada, Y. (1984). Sources of proline-nitrogen in water-stressed soybean (Glycine<br />

max). 11. Fate of '5N-labelled protein. Physiol. Plant., 61: 622-628.<br />

Gage, D. A., Rhodes, D., Nolte, K. D., Hicks, W. A., Leustek, T., Cooper, A. J. L., and Hanson,<br />

A. D. (1997). A new route for synthesis of dimethylsulphoniopropionate in marine algae.<br />

Nature, 387 891-894.<br />

Galili, G. (I 995). Regulation of lysine and threonine synthesis. Plant Cell, 7 899-906.<br />

Galston, A. W. and Sawhney, R. K. (1990). Polyamines in plant physiology. Plant Physiol., 94:<br />

406-4 10.<br />

Gamarnik, A. and Frydman, R. B. (1991). Cadaverine, an essential diamine for the normal root<br />

development of germinating soybean (Glycine max) seeds. Plant Physiol., 97 778-785.<br />

Giddings, T. H. Jr. and Hanson, A. D. (1982). Water stress provokes a generalized increase in<br />

phosphatidylcholine turnover in barley leaves. Planta, 155: 493-501.<br />

Giovanelli, J., Mudd, S. H., and Datko, A. H. (1980). Sulfur amino acids in plants. The Biochernistry<br />

of Plants, Vol. 5 (B. J. Miflin, ed.), Academic Press, New York, pp. 453-505.<br />

Girousse, C., Bournoville, R. and Bonnemain, J.-L. (1996). Water deficit-induced changes in concentrations<br />

in proline and some other amino acids in the phloem sap of alfalfa, Plant Physiol.,<br />

111: 109-113.<br />

Givan, C. V., Joy, K. W., and Kleczkowski, L. A. (1988). A decade of photorespiratory nitrogen<br />

cycling. Trends Biochenz. Sci., 13: 433-437.<br />

Gongalves-Butruille, M., Szajner, P., Torigoi, E., Leite, A., and Arruda, P. (1996). Purification<br />

and characterization of the bifunctional enzyme lysine-ketoglutarate reductase-saccharopine<br />

dehydrogenase from maize. Plant Physiol., 1 IO: 765-771.<br />

Good, A. G. and Crosby, W. L. (1989). Anaerobic induction of alanine aminotransferase in barley<br />

root tissue. Plant Physiol., 90: 1305-1309.<br />

Good, A. G. and Muench, D. G. (1992). Purification and characterization of an anaerobically<br />

induced alanine aminotransferase from barley roots. Plant Physiol., 99: 1520-1525.<br />

Gorham, J. (1995). Betaines in higher plants-biosynthesis and role in stress metabolism. <strong>Amino</strong><br />

<strong>Acid</strong>s and Their Derivatives in Higher Plants (R. M. Wallsgrove, ed.), Society for Experimental<br />

Biology Seminar Series, Vol. 56, Cambridge University Press, Cambridge, pp. 173-<br />

203.<br />

Haglund, B. M. (1980). Proline and valine-cues which stimulate grasshopper herbivory during<br />

drought stress? Nature, 288: 697-698.<br />

Handa, S,, Handa, A. K,, Hasegawa, P. M., and Bressan, R. A. (1986). Proline accumulation and<br />

the adaptation of cultured plant cells to water stress. Plant Physiol., 80: 938-945.<br />

Hanson, A. D. and Hitz, W. D. (1982). Metabolic responses of mesophytes to plant water deficits.<br />

Annu. Rev. Plant Physiol., 33: 163-203,<br />

Hanson, A. D. and Rhodes, D. (1983). I4C tracer evidence for synthesis of choline and betaine via<br />

phosphoryl base inte~ediates in salinized sugarbeet leaves. Plant Physiol., 71: 692-700.<br />

Hanson, A. D., Rathinasabapathi, B., Chamberlin, B., Gage, D. A. (1991). Comparative physiologi-


348 odes et a].<br />

cal evidence that P-alanine betaine and choline-~-su~fate act as compatible osmolytes in<br />

halophytic Lirnoniurn species. Plant Physiol., 97 1199-1205.<br />

Hanson, A. D. and Burnet, M. (1994). Evolution and metabolic engineering of osmoprotectant<br />

accumulation in. higher plants. Cell Biology: Biochemical and Cellular Mechanisms of Stress<br />

Tolerance in Plants (J. H. Cherry, ed.), NATO AS1 Series H, Springer-Verlag, Berlin, pp.<br />

291-302.<br />

Hanson, A. D., Rathinasabapathi, B., Rivoal, J., Burnet, M., Dillon, M. O., and Gage, D. A.<br />

(1994a). Osmoprotecti~e compounds from the Plumbaginaceae: A natural experiment in<br />

metabolic engineering of stress tolerance. Proc. Nutl. Acad. Sci. USA, 91: 306-310.<br />

Hanson, A. D., Rivoal, J., Paquet, L., and Gage, D. A. (1994b). Biosynthesis of 3-dimethylsulfoniopropionate<br />

in ~ol~astonia bifZora (L.) DC: Evidence that S-methylmethionine is an intermediate.<br />

Plant Physiol., 105: 103-1 10.<br />

Hanson, A. D., Rivoal, J., Burnet, M., and Rathinasabapathi, €3. (1995). Biosynthesis of quaternary<br />

ammonium and tertiary sulphonium compounds in response to water deficit. Envjronrnent<br />

and Plant ~etabolisrn: ~lexibili~ and Acclirnation (N. Smirnoff, ed.), Bios Scientific, Oxford,<br />

pp. 189-198.<br />

Hare, P. D. and Cress, W. A. (1997). Metabolic implications of stress-induced proline accumulation<br />

in plants. Plant ~rowth Regul., 21: 79-102.<br />

Hashimoto, T. and Yamada, Y. (1994). Alkaloid biogenesis: Molecular aspects, Annu. Rev. Plant<br />

Physiol. Plant Mol. Biol., 45: 257-285,<br />

Hayser, J. W., De Bruin, D., Kincaid, M. L., Johnson, R. Y., Rodriguez, M. M., and Robinson, N.<br />

J. (1989a). Inhibition of NaC1-induced proline biosynthesis by exogenous proline in halophilic<br />

~~stjchl~s spicata suspension cultures. J. Exp. Bot., 40: 225-232.<br />

Hayser, J. W., Chacon, M. J., and Warren, R. C. (1989b). Characterization of ~-[S-”CC]-proline<br />

biosynthesis in halophytic and nonhalophyticsuspension cultures by I3C NMR. J. Plant<br />

Physiol., 135: 459-466,<br />

He, C.-J,, Finlayson, S. A., Drew, M. C., Jordan, W. R., and Morgan, P. W. (1996a). Ethylene<br />

biosynthesis during aerenchyma fo~ation in roots of maize subjected to mechanical impedance<br />

and hypoxia. Plant Ph~siol., 112: 1679-1685.<br />

He, C.-J., Morgan, P. W., and Drew, M. C. (1996b). Transduction of an ethylene signal is required<br />

for cell death and lysis in the root cortex of maize during aerenchyma formation induced<br />

by hypoxia. Plant Physiol., 112: 463-472.<br />

Hearl, W. G. and Churchich, J. E. (1984). Interactions between 4-aminobutyrate aminotransferase<br />

and succinic semialdehyde dehydrogenase, two mitochondrial enzymes. J. Biol. Chern,, 259<br />

11459-1 1463,<br />

Hitz, W. D., Rhodes, D., and Hanson, A. D, (1981). Radiotracer evidence implicating phosphoryl<br />

and phosphatidyl bases as intermediates in betaine synthesis by water-stressed barley leaves.<br />

Plant Physiol,, 68: 814-822.<br />

HolmstriSm, K.-O., Welin, B., Mandal, A., Krstiansdottir, I., Teeri, T. H., Lamark, T., Str(bm, A.,<br />

and Palva, E. T. (1994). Production of the Escheric~zia coli betaine-aldehyde dehydrogenase,<br />

an enzyme required for the synthesis of the osmoprotectant glycine betaine, in transgenic<br />

plants, Plant J., 6: 749-758.<br />

Hong-qi, Z., Croes, A. F., and Linskens, N. F. (1982). Protein synthesis in germinating pollen of<br />

Petuni~~ Role of proline. Planta, 154: 199-203.<br />

Howden, R., Goldsbrough, P. B., Andersen, C. R., and Cobbett, C. S. (1995a). Cadmium-sensitive,<br />

cad1 mutants of Arab~do~sis thaliana are phytochelatin deficient. Plant Physiol., 107 1059-<br />

1066.<br />

Howden, R., Andersen, C. R., Goldsbrough, P. B., and Cobbett, C. S, (1995b). A cadmium-sensitive,<br />

glutathione-deficient mutant of Ara~ido~sis ~hul~ana. Planr Physiol., 107: 1067-<br />

1073.<br />

Nu, C.-A. A., Delauney, A. J., and Verma, D. P. S. (1992). A bifunctional enzyme (A1-pyrroline-<br />

5-cuboxylate synthetase) catalyzes the first two steps in proline biosynthesis in plants, Proc.<br />

Natl. Acad. Sci. USA, 89: 9354-9358.


Huang, A. H, C. and Cavalieri, A. J. (1979).Prolineoxidase and water-stressinducedproline<br />

accumulation in spinach leaves. Plant Physiol., 63: 531-535.<br />

Ishitani, M., Arakawa, K,, Mizuno, K., Kishitani, S., and Takabe, T. (1993). Betaine aldehyde<br />

d~hydrogenase in the Gramineae: Levels in leaves of both betaine-accumu~ating and nonaccumulating<br />

cereal plants. Plant Cell Physiol., 34: 493-495.<br />

Ishitani, M., Nakamura, T., Han, S. Y., and Takabe, T. (1995). Expression of the betaine aldehyde<br />

dehydrogenase gene in barley in response to osmotic stress and abscisic acid. Plant Mol.<br />

Biol,, 27: 307-315.<br />

James, F., Nolte, K. D., and Hanson, A. D. (1995a). Purification and properties of S-adenosyl-Lmethionine:L-methionine<br />

~-methyltransferase from ~ollaston~a bif2ora leaves. J. Biol.<br />

Chem., 270: 2234422350,<br />

James, F., Paquet, L., Sparace, S. A., Gage, D. A,, and Hanson, A. D, (1995b). Evidence implicating<br />

dimethylsulfoniopropionaldehyde as an intermediate in dimethylsulfoniopropionate biosynthesis.<br />

Plunt Physiol,, 108: 1439-1448.<br />

JimCnez, A,, Hernandez, J. A,, del Rio, L. A., and Sevilla, F. (1997). Evidence for the presence of<br />

the ascorbate-glutathione cycle in mitochondria and peroxisomes of pea leaves. Plant Physi~l.,<br />

114: 275-284.<br />

Jones, M. M., Osmond, C. B., and Turner, N,C. (1980). Accumulation of solutes in leaves of<br />

sorghum and sunflower in response to water deficits. Aust, J. Plant Physiol., 7 193-205.<br />

Karsten, U., Kuck, K., Vogt, C., and Kirst, G. 0. (1996). Dimethylsulf~niopropio~ate production<br />

in phototrophic organisms and its physiological function as a c~oprot~ctant. Biological<br />

Chemistry o ~~MSP and Related ~u~oniu~ Compounds (R. P. Kiene, P. T. Visscher, M. D.<br />

Keller, and G. 0. Kirst, eds.), Plenum Press, New York, pp. 143-153.<br />

Katz, A. and Tal, M. (1980). Salt tolerance in the wild relatives of cultivated tomato: Proline<br />

accumulation in callus tissue of Lycopersicon esculentum and L. per~~ianum. Z. Pflanzenphysiol.,<br />

98: 429-435.<br />

Kende, H. (1993). Ethylene biosynthe~is. Annu. Rev, Plant. Physiol., 44: 283-307.<br />

Kishitani, S., Watanabe, K., Yasuda, S., Arakawa, K., and Takabe, T, (1994). Accumulation of<br />

glycinebetaine during cold acclimation and freezing tolerance in leaves of winter and spring<br />

barley plants. Plant Cell Environ., 17 89-95.<br />

Kishor, K. P. B,, Hong, Z., Miao, 6. H., Hu, C.-A. A., and Verma, D. P. S. (1995). Overexpression<br />

of ~'-py~oline-5-carboxylate synthetase increases proline production and confers osmotolerance<br />

in transgenic plants. Plant Physiol., 108: 1387-1394.<br />

Kiyosue, T., Yoshiba, Y., Yamaguchi-Shinozaki, IS., and Shinozaki, K. (1996). A nuclear gene,<br />

encoding mitochondrial proline dehydrogenase, an enzyme involved in proline metabolism,<br />

is upregulated by proline but downregulated by dehydration in Arabido~sis. Plant Cell, 8:<br />

1323-1335.<br />

Klapheck, S., Schlunz, S., and Bergmann, L. (1995). Synthesis of phytochelatins and homophytochelatins<br />

in Pisum sativum L. Plant Physid., 107: 515-521.<br />

Kleczkowski, L. A, and Givan, C. V. (1988). Serine formation in leaves by mechanisms other than<br />

the glycolate pathway. J. Plant Physiol., 132: 641-652.<br />

Kreuz, K., Tommasini, R., and Martinoia, E. (1996). Old enzymes for a new job: Herbicide detoxification<br />

in plants. Plant Physiol., 111: 349-353,<br />

Kumar, A,, Taylor, M., Mad Arif, S, A,, and Davies, H. (1996). Potato plants expressing antisense<br />

and sense S-adenosylmethionine decarboxylase (SAMDG) transgenes show altered levels<br />

of polyaminesandethylene:Antisenseplantsdisplayabnormalphenotypes. Plant J., 9:<br />

147-158.<br />

Lamark, T., Kaasen, I., Eshoo, M. W., Falkenberg, P., McDougall, J,, and Strpfm, A. R. (1991).<br />

DNA sequence and analysis of the bet genes encoding the osmoregulatory choline-glycine<br />

betaine pathway of Escherichia coli. Mol. ~icrobiol~, 5: 1049-1064.<br />

Lamark, T., Rokenes, T. P., McDougall, J., and Strpfm, A. R, (1996). The complex bet promoters<br />

of Escherichiu coli: Regulation by oxygen (ArcA), choline (BetZ), and osmotic stress. J.<br />

Bacteriol., 178 1655-1662.


3<br />

Lansac, A. R., Sullivan, C, Y., and Johnson, B. E. (1996). Accumulation of free proline in sorghum<br />

(Sorghum bicolor) pollen. Can. J. Bot., 74: 40-45.<br />

Lerma, C., Hanson, A. D., and Rhodes, D. (1988). Oxygen-1 8 and deuterium labeling studies of<br />

choline oxidation by spinach and sugar beet. Plant Physiol., 88: 695-702.<br />

Lerma, C., Rich, P. J., Ju, G. C., Yang, W.-J., Hanson, A. D., and Rhodes, D. (1991). Betaine<br />

deficiency in maize: Complementation tests and metabolic basis. Plant Physiol., 95: 11 13-<br />

1119.<br />

Li, Z.-S., Zhao, Y., and Rea, P. A. (1995). Magnesium adenosine 5’-t~phosphate-energized transport<br />

of glutathione- conjugates by plant vacuolar membrane vesicles. Plant Physiol., 107<br />

1257-1 268.<br />

Lieberman, M. (1979). Biosynthesis and action of ethylene. Annu. Rev. Plant Physiol., 30: 533-<br />

591.<br />

Lilius, G., Holmberg, N., and Biilow, L. (1996). EnhancedNaClstresstoleranceintransgenic<br />

tobacco expressing bacterial choline dehydrogenase. Biotechnology, 14: 177-1 80.<br />

Ling, V., Snedden, W. A., Shelp, B. J., and Assmann, S. M. (1994). Analysis of a soluble calmodulin<br />

binding proteinfromfavabeanroots:Identificationofglutamate decarboxylase as a<br />

calmodulin-activated enzyme. Plant Cell, 6: 1135-1 143.<br />

Liu, J. and Zhu, J.-K. (1997). Proline accumulation and salt-stress-induced gene expression in a<br />

salt-hypersensitive mutant of Arabi~o~s~s. Plant Physiol,, 114: 591-596.<br />

Low, P. S. (1985). Molecular basis of the biological compatibility of nature’s osmolytes. <strong>Transport</strong><br />

Processes, lono- and ~smoregulation (R. Gilles and M. Gilles-Baillien, eds.), Springer-Verlag,<br />

Berlin, pp. 469-477.<br />

Maitani, T., Kubota, H., Sato, K., and Yamada, T. (1996). The composition of metals bound to<br />

class 111 metallothionein (phytochelatin and its desglycyl peptide) induced by various metals<br />

in root cultures of Rubia tinctorum. Plant Physiol., 110: 1145-1 150.<br />

Malmberg, R. L., Smith, K.E., Bell, E., and Cellino, M.L. (1992). Arginine decarboxylase of<br />

oats is clipped from a precursor into two polypeptides found in the soluble enzyme. Plant<br />

Physiol., 100: 146-152.<br />

Marrs, I(. A. (1996). The function and regulation of glutathione S-transferases in plants. Annu.<br />

Rev. Plant Physiol. Plant Mol. Bioi., 47: 127-157.<br />

Martin, M. N., Cohen, J. D., and Saftner, R. A. (1995). A new l-aminocyclopropan~-l-carboxyl~c<br />

acid-conjugating activity in tomato fruit. Plant Physiol., 109: 917-926.<br />

Martin, M. N. and Saftner, R. A. (1995). Purification and characterization of I-aminocyclopropane-<br />

1-carboxylic acid ~-malonyltransferase from tomato fruit. Plant Physiol,, 108: 1241-1249.<br />

Mayer, R. R., Cherry, J. H., and Rhodes, D. (1990). Effects of heat shock on amino acid metabolism<br />

of cowpea cells. Plant Physiol., 94: 796-810.<br />

McCue,K. F. and Hanson, A. D. (1992). Salt-inducible betaine aldehyde dehydrogenase from<br />

sugar beet: cDNA cloning and expression. Plant Mol. Bioi., 18: 1-1 1.<br />

McLean, W. F. H., Blunden, G., and Jewers, K. (1996). Quaternary ammonium compounds in the<br />

Capparaceae. Biochem. Syst. Ecol., 24: 427-434.<br />

Meister, A. (1983). Selective modification of glutathione metabolism. Science, 220: 472-477.<br />

Miyazaki, J. H. and Yang, S. F. (1.987). The methionine salvage pathway in relation to ethylene<br />

and polyamine biosynthesis. Physiol. Plant., 69: 366-370.<br />

Morris, C. J., Thompson, J. F., and Johnson, C. M. (1969). Metabolism of glutamic acid and N-<br />

acetylglutamic acid in leaf discs and cell-free extracts of higher plants. Plant Physiol., 44:<br />

1023-1026.<br />

Mudd, S. H. and Datko, A, H. (1989a). Synthesis of methylated ethanolamine moieties. Regulation<br />

by choline in Lemna. Plant Physiol., 90: 296-305.<br />

Mudd, S. H. and Datko, A. H. (1989b). Synthesis of methylated ethano~amine moieties. Regulation<br />

by choline in soybean and carrot. Plant Physiol., 90: 306-310.<br />

Naidu, B. P., Jones, G. P., Paleg, L. G., and Poljakoff-Mayber, A. (1987). Proline analogues in<br />

Mela~euca species: Response of Melaleuca lanceolata and M. uncinata to water stress and<br />

salinity. Aust. J. Plant Physiol., 14: 669-677.


Naidu B. P., Paleg, L. G., and Jones, G. P. (1992). Nitrogenous compatible solutes in droughtstressed<br />

Medicago spp. Phytochem,istry, 31: 1195-1 197.<br />

Nakamura, T., Yokota, S., Muramoto, Y., Tsutsui, K., Oguri, Y., Fukui, K., and Takabe, T. (1997).<br />

Expression of a betaine aldehyde dehydrogenase gene in rice, a glycinebetaine nonaccumulator,<br />

and possible localization of its protein in peroxisomes. Plunt J., 11: 11 15-1 120.<br />

Nishiguchi, M. K. and Somero, G. N. (1992). Temperature- and concentration-dependence of com-<br />

~atibi~ity of the organic osmolyte ~-~imethy~sulfoniopropionate. Cryobiology, 29: 118-1 24.<br />

Nolte, K. D., Hanson, A. D., and Gage, D. A. (1997). Proline accumulation and ethyla at ion to<br />

proline betaine in Citrus: ~mplicat~ons for genetic engineering of stress resistance. J. Am.<br />

SOC. Hort. Sci., 122: 8-1 3.<br />

Nomura, M., Ishitani, M., Takabe, T., Rai, A. K., and Takabe, T. (1995a). Synechococcus sp.<br />

PCC7942 transformed with Escherichia coli bet genes produces glycine betaine from choline<br />

and acquires resistance to salt stress. Plant Physiol., 107: 703-708.<br />

Nomura, M., Muramoto, Y., Yasuda, S., Takabe, T., and Kishitani, S. (1995b). The accumulation<br />

of glycinebetaine during cold acclimation in early and late cultivars of barley. ~uph~tica,<br />

83: 247-250.<br />

Nonami, H., Wu, Y., and Boyer, J. S. (1997). Decreases growth-induced water potential: A primary<br />

cause of growth inhibition at low water potentials. Plant Physiol., 114: 501-509.<br />

Oaks, A., Mitchell, D.J., Barnard, R, A., and Johnson, F. J. (1970). The regulation of proline<br />

biosynthesis in maize roots. Can. J. Bot., 48: 2249-2258.<br />

Ober, E. S. and Sharp, R. E. (1994). Proline accu~ulation in maize (Zea mays L.) primary roots<br />

at low water potentials. I. Requirement for increased levels of abscisic acid. Plant P~~~sjo~.,<br />

105: 981-987.<br />

Anntt. Rev.<br />

Oliver, D. J, (1994). The glycine decarboxylase complex from plant mitochondria.<br />

Plant Physiol. Plant Mol. Bioi., 45: 323-337.<br />

Paquet, L., Rathinasabapathi, B., Saini, H., Zamir, L., Gage, D. A., Huang, Z.-H., and Hanson, A.<br />

D. (1994). Accumulation of the compatible solute 3-dimethy~sulfoniopropionate in sugarcane<br />

and its relatives, but not other gramineous crops. Aust. J. Plant Physiol., 21: 37-48.<br />

Pastori, G. M. and Trippi, V. S. (1992). Oxidative stress induces high rate of glutathione reductase<br />

synthesis in a drought-resistant maize strain. Plant Cell Physiol., 33: 957-961.<br />

Patterson, B. D. and Graham, D. (1987). Temperature and metabolism. The biochemist^ o~~lunts,<br />

Vol. 12 (D. D. Davies, ed.), Academic Press, New York, pp. 153-199.<br />

Peng, Z., Lu, Q., and Verma, D. P, S. (1996). Reciprocal regulation of A'-py~oline-5-carboxylate<br />

synthetase and proline dehydrogenase genes control proline levels during and after osmotic<br />

stress in plants, Mol. Gen. Genet,, 253: 334-341.<br />

Perez-Amador, M. A., Carbonell, J., Navarro, J. L., Moritz, T., Beale, M. H., Lewis, M. J., and<br />

Hedden, P. (1996). N4-Hexanoylspermidine, a new polyamine-re~a~ed compound that accumulates<br />

during ovary and petal senescence in pea, Plant Physiol., 110: 1177-1 186.<br />

Pickett, C. G. and Lu, A. Y. H. (1989). Glutathione S-transferases: Gene structure, regulation, and<br />

biological function. Annu. Rev. Biochem., 58: 743-764.<br />

Prabhu, V., Chatson, K. B., Abrams, G. D., and King, J. (1996). I3C Nuclear magnetic resonance<br />

detection of interactions of serine hydroxymethyltransferase with C~-tetrahydrofolate synthase<br />

and glycine decarboxylase complex activities in Arabjdopsis. Plant Physiol., 112: 207-<br />

216,<br />

Ramputh, A.4. and Bown, A. W. (1996). Rapid y-aminobutyric acid synthesis and the inhibition<br />

of the growth and development of oblique-banded leaf-roller larvae. Plant Physiol., 111:<br />

1349-1 352.<br />

Rastogi, R., Dulson, J,, and Rothstein, S. J. (1993). Cloning of tomato (Lycopersicon esculentu~z<br />

ill.) arginine decraboxylase gene and its expression during fruit ripening. Plant Physiol.,<br />

103: 829-834,<br />

Ratcliffe, R. 6, (1995). Metabolic aspects of the anoxic response in plant tissue. ~nvironment and<br />

Plant M~tabolis~: ~le~ibili~ and Acclimation (N. Smirnoff, ed.), Bios Scientific, Oxford,<br />

pp, 111-127.


Rathinasabapathi, B., McCue, K. F., Gage, D. A,, and Hanson, A. D. (1994). Metabolic engineering<br />

of glycine betaine biosynthesis: Plant betaine aldehyde dehydrogenases lacking typical<br />

transit peptides are targeted to tobacco chloroplasts where they confer betaine aldehyde<br />

resistance. Plant, 193: 155-1 62,<br />

Rathinasabapathi, B., Burnet, M., Russell, €3. L., Gage, D. A., Liao, P.-C,, Nye, G. J,, Scott, P,,<br />

Golbeck, J, W,, and Wanson, A. D, (1997). Choline monooxygenase, an unusual iron-sulfur<br />

enzymecatalyzingthefirststepofglycinebetainesynthesisinplants:Prostheticgroup<br />

characterizat~on and cDNA cloning. Proc. Natl. Acad. Sci. USA, 94: 3454-3458.<br />

Rauser, W, E. (1990). Phytochelatins. Annu. Rev. Bioche~., 59 61-86.<br />

Rauser, W. E. (1995). ~hytochelatins and related peptides. Structure, biosynthesis, and function.<br />

Plant Physiol., 109 1141-1 149.<br />

Rauser, W. E. and Meuwly, P, (1995). Retention of cadmium in roots of maize seedlings. Role of<br />

complexation by phytochelatins and related thiol peptides. Plant Physiol., 109: 195-202.<br />

Reese, R. N. and Wagner, C. J. (1987). Effects of buthionine sulfoximine on Cd-binding peptide<br />

levels in suspension-cultured tobacco cells treated with Cd, Zn, or Cu. Plant Physlol., 84:<br />

574-577.<br />

Rentsch, D,, Hirner, B., Schmeizer, E., and Frommer, W. €3. (1996). Salt stress-induced proline<br />

transporters and salt stress-repressed broad specificity amino acid permeases identified by<br />

suppression of a yeast amino acid permease-targeting mutant. Plant Cell, 8: 1437-1446.<br />

Rhodes, D., Handa, S., and Bressan, R. A. (1986). Metabolic changes associated with adaptation<br />

of plant cells to water stress. Plant Physiol., 82: 890-903.<br />

Rhodes, D. (1987). Metabolic responses to stress. The Bioche~is~~ of Plants,Vol. 12 (D. D. Davies,<br />

ed.), Academic Press, New York, pp. 201-241.<br />

Rhodes, D. and Handa, S. (1989). <strong>Amino</strong> acid metabolism in relation to osmotic adjustment in<br />

plant cells. ~nviron~ental Stress in Plants: Bioche~ic~l and P~ysiolog~cal ~echan~sms (J.<br />

H. Cherry, ed.), NATO AS1 Series, Vol. G19, Springer-Verlag, Berlin, pp. 41-62.<br />

Rhodes, D, and Hanson, A. D. (1993). Quaternary ammonium and tertiary sulfonium compounds<br />

in higher plants. Annu. Rev. Plant Physio~, Plant Mol, Biol,, 44: 357-384.<br />

Rhodes, D., Yang, W.-J., Samaras, Y., Wood, I


~ Stewart,<br />

D. (1995). Proline accumulation during drought and salinity. Environment and Plant Met~bolism:<br />

Flexibili~ and Acclimation (N. Smirnoff, ed.), Bios Scientific, Oxford, pp. 161-<br />

18’7.<br />

Saneoka, H., Nagasaka, C,, Hahn, D. T., Yang, W.-J., Premachandra, G. S,, Joly, R. J,, and Rhodes,<br />

D. (1995). Salt tolerance of glycinebetaine-deficient and -containing maize lines. Plant<br />

Physiol,, 107: 631-4338.<br />

SavourtS, A., Hua, X,-J., Bertauche, N., Van Montagu, M., and Verbmggen, N, (1997). Abscisic<br />

acid-independent and abscisic acid-dependent regulation of proline biosynthesis following<br />

cold and osmotic stresses in A~ffbido~sis thffliana. Mol. Gen. Genet., 254: 104-109.<br />

Sells, G. D. and Koeppe, D. E. (1981). Oxidation of proline of mitochondria isolated from waterstressed<br />

maize shoots. Plant Physiol., 68: 1058-1063.<br />

Sharp, R. E., Boyer, J. S., Nguyen, H. T., and Hsiao, T. C. (1996). Genetically engineered plants<br />

resistant to soil drying and salt stress: How to interpret osmotic relations. Plant Physiol.,<br />

110: 1051-1053.<br />

Sharp, R. E., Wu, Y., Voetberg, G. S,, Saab, 1. N., and LeNoble, M. E. (1994). Confi~ation that<br />

abscisic acid accumulation is required for maize primary root elongation at low water potentials.<br />

J. Exp. Bot,, 45: 1743-1751,<br />

Shelp, B. J., Walton, C. S,, Snedden, W. A., Tuin, L, G,, Oresnik, I. J., and Layzell, D. €3, (1995).<br />

GABA shunt in developing soybean seed is associated with hypoxia. Physiol. Plant, 94:<br />

2 19-228,<br />

Shen, B., Jensen, R. G.,andBohnert,H.J. (1997).Increasedresistancetooxidativestressin<br />

transgenic plants by targeting mannitol biosynthesis to chloroplasts. Plant Physiol., 113:<br />

1177-1 183.<br />

Smirnoff, N. and Cumbes, Q. J. (1989). Hydroxyl radical scavenging activity of compatible solutes.<br />

Phytochemistry, 28: 105’7-1060.<br />

Smirnoff, N. (1993). The role of active oxygen in the response of plants to water deficit and<br />

desiccation. New Phytol., 125: 27-58.<br />

Smimoff, N. (1995). Antioxidant systems and plant response to the environment. Envi~o~ment and<br />

Plant ~etabolism: Flexibili~ and Acclimation (N. Smirnoff, ed,), Bios Scientific, Oxford,<br />

pp. 2 17-243.<br />

Smith, T. A., Bagni, N., and Fracassini, D, S. (1979). The formation of amines and their derivatives<br />

in plants. ~itroge~ Assimilation of Plants (E, J. Hewitt and C. V. Cutting, eds.), Academic<br />

Press, New York, pp. 557-570.<br />

Snedden, W. A., Arazi, T,, Fromm, H., and Shelp, B. J. (1995). Calciu~calmodulin activation of<br />

soybean glutamate decarboxylase. Plant Physiol., 108 543-549.<br />

Somero, G. N. (1986). Protons, osmolytes, and fitness of internal milieu for protein function. Am.<br />

J. Physiol., 251: R197-R213.<br />

Steffens, J. C.,Hunt, D. F.,andWilliams, €3. G. (1986).Accumulationofnon-proteinmetalbinding<br />

polypeptides (y-glutamyl-cysteinyl),,-glycine in selected cadmium resistant tomato<br />

cells. J. Biol. Chem., 261: 13879-13882.<br />

Stewart, C. R. (1981). Proline accumulation: biochemical aspects. The P~zysiology and Biochemistry<br />

of ~~o~ght Resistance in Plants (L. G. Paleg and D. Aspinall, eds.), Academic Press,<br />

Sydney, pp. 243-259.<br />

C. R. and Boggess, S. F. (1978). Metabolism of [5FH]proline by barley leaves and its use<br />

in measuring the effects of water stress on proline oxidation. Plant Physiol., 61: 656657.<br />

Stewart, C. R,, Boggess, S. F,, Aspinall, D., and Paleg, G. (1977). Inhibition of proline oxidation<br />

by water stress. Plant Ph~lsiol., 59: 930-932.<br />

Stewart, G. R. and Lee, J. A, (1974). The role of proline accumulation in halophytes. Planta, 120:<br />

279-289,<br />

Stewart, G. R. and Larher, I;. (1980).Accumulationofaminoacidsandrelatedcompoundsin<br />

relationtoenvironmentalstress. The biochemist^ of Plants. Vol. 5 (B. J. Miflin,ed.),


3 s et al.<br />

Streeter, J. G. and Thompson, J. F. (1972). Anaerobic accumulation of y-aminobutyric acid and<br />

alanine in radish leaves (Raphanus sativus L.). Plant Physiol., 49: 572-578.<br />

Strgm, A. R., Le Ruddier, D., Jakowec, M.W., Bunnell, R. C., and Valentine, R. C, (1983).<br />

Osmoregulatory (Osm) genes and osmoprotective compounds. Genetic Engineering of<br />

Plants. An Agricultural Perspective (T. Kosuge, C. P. Meredith, and A. Hollaender, eds.),<br />

Plenum Press, New York, pp. 39-59.<br />

Summers, J. E., Voesenek, L. A. C. J,, Blom, C. W. P. M., Lewis, M. J., and Jackson, M. B.<br />

(1996). Potamogeton pectinatus is constitutively incapable of synthesizing ethylene and<br />

lacks 1 -aminocyclopropane- 1-carboxylic acid oxidase. Plant Physiol,, I1 I: 901-908.<br />

Thomas, J. C., De Armond, R. L., and Bohnert, H. J. (1992). Influence of NaCl on growth, proline,<br />

and phosphoenolpyruvate carboxylase levels in ~esemb~anthemum crystallin~m suspension<br />

cultures. Plant Physiol., 98: 626-63 1.<br />

Thompson, J. F. (1980). Arginine synthesis, proline synthesis, and related processes, The Biochemistry<br />

of Plants, Vol. 5 (B. J. Miflin, ed.), Academic Press, New York, pp. 375-402.<br />

Tirnasheff, S. N, (1993). The control of protein stability and association by weak interactions with<br />

water: How do solvents affect these processes? Annu. Rev. Biophys. Biornol. Struct., 22:<br />

67-97.<br />

Torsethaugen, G., Pitcher, L. H., Zilinskas, B. A., Pell, E. J. (1997). Overproduction of ascorbate<br />

peroxidase in the tobacco chloroplast does not provide protection against ozone. Plant Physiol.,<br />

114: 529-537.<br />

Treichel, S. (1975). The effect of NaCl on the concentration of proline in different halophytes. Z.<br />

P~an~enphysjol., 76: 56-68.<br />

Treichel, S. (1986). The influence of NaCl on ~'-pyrroline-5-carboxylate reductaseinprolineaccumulating<br />

cell suspension cultures of ~ese~nb~~anthemum ~od~orum and other halophytes.<br />

Plant Physiol. 67 173-181.<br />

Trossat, C., Nolte, K. D., and Hanson, A. D. (1996). Evidence that the pathway of dimethylsulfoniopropionate<br />

biosynthesis begins in the cytosol and ends in the chloroplast. Plant Physiol.,<br />

I1 1: 965-973.<br />

Trossat, C., ~athinasabapathi, B., and Hanson, A. I). (1997). Transgenically expressed betaine<br />

aldehyde dehydrogenase efficiently catalyzes oxidation of dimethylsulfoniopropionaldehyde<br />

and ~arninoaldehydes. Plant Physiol., 113: 1457-1461.<br />

Turano, F. J., Thakkar,.S. S., Fang, T., and Weisemann, J. M. (1997). ~haracterization and expression<br />

of NAD(H)-dependent glutamate dehydrogenase genes in Arabi~opsis. Plant Physiol.,<br />

113: 1329-1341.<br />

Ulmasov, T., Ohmiya, A., Hagen, G.,and Guilfoyle, T. (1995). The soybean GH24 gene that<br />

encodes a glutathione S-transferase has a promoter that is activated by a wide range of<br />

chemical agents. Plant Physiol., 108: 919-927.<br />

Vandewalle, I. and Olsson, R. (1983). The ~-aminobutyric acid shunt in ge~inating Sinapis alba<br />

seeds. Plant Sci. Lett., 31: 269-273.<br />

Verbruggen, N., Villarroel, R., and Van Montagu, M. (1993). Osmoregulation of a pyrroline-5-<br />

carboxylate reductase gene in Arabidopsis thaliana, Plant Physiol., 103: 771-781.<br />

Visser, E. J. W., Gohen, J. D., Barendse, G. W. M., Mom, C. W. P. M., and Voesenek, L. A. C.<br />

J. (1996). An ethylene-mediated increase in sensitivity to auxin adventitious root formation<br />

in flooded Rumex palustris Sm. Plant Physiol., 112: 1687-1692.<br />

Voetberg, G. S. and Sharp, R. E. (1991). Growth of the maize primary root at low water potentials.<br />

111. Role of increased proline deposition in osmotic adjustment. Plant Physiol., 96: 1125-<br />

1130.<br />

VojtechovB, M., Hanson, A. D., and MuRoz-Clares, R. A. (1997). Betaine-aldehyde dehydrogenase<br />

from amaranth leaves efficiently catalyzes the NAD-dependent oxidation of dimethylsulfoniopropionaldehyde<br />

to dimethylsulfonioprop~onate. Arch. Biochem. Biophys., 337: 81-88.<br />

Walden, R., Cordeiro, A., and Tiburcio, A. F. (1997). Polyamines: small molecules triggering<br />

pathways in growth and development. Plant Physiol., 113: 1009-1013.


Wallace, W., Secor, J., and Schrader, L. E. (1984). Rapid accumulation of ~-aminobutyric acid and<br />

alanine in soybean leaves in response to an abrupt transfer to lower temperature, darkness, or<br />

mechanical manipulation. Plant Physiol., 75: 170-175.<br />

Watson, M. B. and Malmberg, R. L. (1996). Regulation of Arabjdopsis tkaliana (L.) Heynh arginine<br />

decarboxylase by potassium deficiency stress. Plant Physiol., 111: 1077-1083.<br />

Weigel, P., Weretilnyk, E. A., and Hanson, A. D. (1986). Betaine aldehyde oxidation by spinach<br />

chloroplasts. Plant Physiol. 82: 753-759.<br />

Weigel, P., Lerma, C., and Hanson, A. D. (1988). Choline oxidation by intact spinach chloroplasts.<br />

Plant Physiol., 86 54-60.<br />

Wereti~nyk, E. A. and Hanson, A. D. (1988). Betaine aldehyde dehydrogenase polymo~hism in<br />

spinach: genetic and biochemical characterization. Biockem. Genet., 26 143-15 1.<br />

Weretilnyk, E. A. and Hanson, A. D. (1989). Betaine aldehyde dehydrogenase from spinach leaves:<br />

purification, in vitro translation of the mRNA, and regulation by salinity. Arch. Biochem.<br />

Biophys., 271: 56-63.<br />

Weretilnyk, E. A., Bednarek, S., McCue, K. F., Rhodes, D., Hanson, A. D. (1989). Comparative<br />

biochemical and immunological studies of the glycine betaine synthesis pathway in diverse<br />

families of dicotyledons, Planta, 178: 342-352.<br />

Weretilnyk, E. A, and Hanson, A. D. (1990). Molecular cloning of a plant betaine-aldehyde dehydrogenase,<br />

an enzyme implicated in adaptation to salinity and drought. Proc. Natl. Acud.<br />

Sci. USA, 87 2745-2749.<br />

Weretilnyk, E. A. and Summers, P. S. (1992). Betaine and choline metabolism in higher plants.<br />

Biosynthesis and Molecular Regulation of <strong>Amino</strong> <strong>Acid</strong>s in Plants (B. K. Singh, H. E. mores,<br />

and J. C, Shannon, eds.), American Society of Plant Physiologists, Waverly Press, Baltimore,<br />

pp. 89-97.<br />

Weretilnyk, E. A., Smith, D, D., Wilch, G. A., and Summers, P. S. (1995). Enzymes of choline<br />

synthesis in spinach. Response of phospho-base N-methyltransferase activities to light and<br />

salinity. Plant Physiol., 109: 1085-1091.<br />

Williamson, C. L. and Slocum, R. D. (1992). Molecular cloning and evidence for osmoregulation<br />

of the A'-pyrroline-5-carboxylate @roc) gene in pea (Pisum sativum L.). Plant Physiol.,<br />

100: 1464-1 470,<br />

Wolfe, G. V,, Steinke, M., and Kirst, G. 0. (1997). Grazing-activated chemical defence in a unicellular<br />

marine alga. Nature, 387 894-897.<br />

Wood, K. V., Stringham, K. J., Smith, D. L., Volenec, J. J., Hendershot, K. L., Jackson, K. A.,<br />

Rich, P. J., Yang, W.-J., and Rhodes, D. (1991). Betaines of alfalfa: Characterization by<br />

fast atom bombardment and desorption chemical ionization mass spectrometry. Plant Physiol.,<br />

96: 892-897.<br />

Wyn Jones, R. G., Storey, R., Leigh, R. A,, Ahmad, N., and Pollard, A. (1977). A hypothesis on<br />

cytoplasmic osmoregulation. Regulation of Cell ~~rnbrane Activities in Plants (E. Marri:<br />

and 0. Cifferi, eds.), Elsevier, Amsterdam, pp. 121-136.<br />

Wyn Jones, R. G. and Storey, R. (1981). Betaines. The Physiology and biochemist^ of Drougkt<br />

Resistance in Plants (L. G. Paleg and D. Aspinall, eds.), Academic Press, Sydney, pp.<br />

171-204.<br />

Yancey, P. H., Clark, M. E., Hand, S. C., Bowlus, R. D., and Somero, G. N. (1982). Living with<br />

water stress: Evolution of oxmolyte systems. Science, 217 1214-1222.<br />

Yancey, P. H. (1994). Compatible and counteracting solutes. Cellular and ~oleculur Physiology<br />

of Cell Volume Regulation (K. Strange, ed.), CRC Press, Boca Raton, FL,, pp, 81-109.<br />

Yang, G., Rhodes, D., and Joly, R. J. (1996). Effects of high temperature on membrane stability<br />

and chlorophyll fluorescence in glycinebetaine-deficient and glycinebetaine-containing<br />

maize lines. Aust, J. Plant Physiol., 23: 437443.<br />

Yang, S. F. and Hoffman, N. E. (1984). Ethylene biosynthesis and its regulation in higher plants.<br />

Annu. Rev. Plant Physiol., 35: 155-189,<br />

Yang, S. F. (1989). Metabolism of I ~aminocyclopropane-1-carboxylic acid in relation to ethylene


iosynthesis, Plant ~ ~~~ogen ~etabolis~<br />

(J. E, Poulton, J. T. Romeo, and E. E. Conn, eds.).<br />

Rec. Adv, Phyt~~he~., 23: pp. 263-287.<br />

Yang, W.-J., Nadolska-Orczyk, A., Wood, K. V,, Wahn, D, T,, Rich, P, J., Wood, A. J., Saneoka,<br />

H., Premachandra, G. S,, Bonham, C. C., Rhodes, J. C., Joly, R. J,, Samaras, Y., Goldsbrough,<br />

P. B., and Rhodes, D. (1995). Near-isogenic lines of maize differing for glycinebetaine.<br />

Plant Phys~ol., 107 621430.<br />

Yoshiba, Y,, Kiyosue, T., Katagiri, T., Ueda, H,, ~izoguchi, T., Yamaguchi-Shinozaki, K., Wada,<br />

K., Narada, Y., and Shinozaki, K. (1995). Correlation between the induction of a gene for<br />

A1-pyrro1ine-5-carboxylate sy~thetase and the accumu~ation of proline in Arab~dopsis thaliana<br />

under osmotic stress, Plant J,, 7 751-760.

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

Saved successfully!

Ooh no, something went wrong!