estimation of metabolic fluxes

Upload: liderrojo88

Post on 04-Jun-2018

217 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/14/2019 Estimation of Metabolic Fluxes

    1/12

    TECHNICAL ADVANCE

    Estimation of metabolic fluxes, expression levels and meta-bolite dynamics of a secondary metabolic pathway in potato

    using label pulse-feeding experiments combined with kineticnetwork modelling and simulation

    Elmar Heinzle1,2,*, Fumio Matsuda3, Hisashi Miyagawa2,3, Kyo Wakasa3,4 and Takaaki Nishioka2

    1Biochemical Engineering Institute, Saarland University, D-66123 Saarbrucken, Germany,2Division of Applied Life Sciences, Agricultural Department, Kyoto University, Kyoto 606-8502, Japan,3Plant Functions and Their Control, CREST, Japan Science and Technology Agency, 3-4-5 Nihonbashi, Chuo, Tokyo 103-0027,

    Japan, and4Department of Agriculture, Tokyo University of Agriculture, 1737 Funako, Atsugi, Kanagawa 243-0034, Japan

    Received 16 July 2006; revised 13 November 2006; accepted 23 November 2006.*For correspondence (fax +49 681 302 4572; e-mail [email protected]).

    Summary

    In this paper we present a method that allows dynamic flux analysis without a priorikinetic knowledge. This

    method was developed and validated using the pulse-feeding experimental data obtained in our previous

    study (Matsuda et al., 2005), in which incorporation of exogenously applied L-phenylalanine-d5 into seven

    phenylpropanoid metabolites in potato tubers was determined. After identification of the topology of the

    metabolic network of these biosynthetic pathways, the system was described by dynamic mass balances in

    combination with power-law kinetics. After the first simulations, some reactions were removed from the

    network because they were not contributing significantly to network behaviour. As a next step, the exponents

    of the power-law kinetics were identified and then kept at fixed values during further analysis. The model was

    tested for statistical reliability using Monte Carlo simulations. Most fluxes could be identified with highaccuracy. The two test cases, control and after elicitation, were clearly distinguished, and with elicitation

    fluxes toN-p-coumaroyloctopamine (pCO) andN-p-coumaroyltyramine (pCT) increased significantly, whereas

    those for chlorogenic acid (CGA) and p-coumaroylshikimate decreased significantly. According to the model,

    increases in the first two fluxes were caused by induction/derepression mechanisms. The decreases in the

    latter two fluxes were caused by decreased concentrations of their substrates, which in turn were caused by

    increased activity of the pCO- and pCT-producing enzymes. Flux-control analysis showed that, in most cases,

    flux control was changed after application of elicitor. Thus the results revealed potential targets for improving

    actions against tissue wounding and pathogen attack.

    Keywords: metabolic flux analysis, metabolic control analysis, pulse-feeding experiment, kinetic network

    modelling, phenylpropanoid pathway, potato.

    Introduction

    Plants, as major primary producers of biochemical sub-

    stances, have enormous biosynthetic potential that could be

    explored more fully using methods of metabolic engineer-

    ing, as indicated by a series of recent reviews (Hanson and

    Shanks, 2002; Rontein et al., 2002). The basis for rational

    design of metabolic paths is a thorough quantitative

    understanding of the paths, including an understanding of

    their kinetics and their regulation at the levels of genes,

    proteins and metabolites (Bailey, 1998; Giersch, 2000; Niel-

    sen, 1998; Stephanopoulos et al., 1998). Metabolic flux

    176 2007 The AuthorsJournal compilation 2007 Blackwell Publishing Ltd

    The Plant Journal (2007) 50, 176187 doi: 10.1111/j.1365-313X.2007.03037.x

  • 8/14/2019 Estimation of Metabolic Fluxes

    2/12

    analysis is a major tool in this field, providing quantitative

    data about reaction rates. Relative to their importance, such

    quantitative analyses have notbeen applied widely in plants.

    This is probably due to the fact that the wide range of

    methods developed for microbial systems and also mam-

    malian cells in recent years (Dauner and Sauer, 2000; Hel-

    lerstein, 2004; Hellerstein and Neese, 1999; Wahlet al., 2004;

    Wiechert, 2001; Wiechertet al., 2001; Wittmann, 2002; Witt-

    mann and Heinzle, 1999, 2001a,b) have been adapted only

    partially for plant systems (Ratcliffe and Shachar-Hill, 2006;

    Roscher et al., 2000). Several reasons can be identified to

    explain this: (i) plant genetics are more complex than

    microbial genetics, and the Arabidopsis and rice genomes

    have been sequenced in their entirety only relatively re-

    cently; (ii) experimentation with plants is generally more

    complex and less suited to well defined laboratory experi-

    ments compared with microbial systems; (iii) under most

    relevant photosynthetic conditions, plants assimilate carbon

    dioxide, which is not well suited for classical metabolic flux

    analysis methods using 13

    C-label distribution measure-ments; (iv) plants are usually highly compartmented, which

    complicates metabolic flux analysis; and (v) it is difficult to

    carry out experiments in plants under steady-state condi-

    tions (such as those used in the chemostat technique for

    microbial cultures), which are desirable for metabolic flux

    analysis methods based on well-established metabolic

    stoichiometry.

    While quantitative flux analyses of central pathways and

    polymeric reactions using13C-labelled precursor substances

    in plantsundersteady-stateconditionshave been carried out

    quite extensively in recent years (Glawischniget al., 2002;

    Hellersteinand Neese, 1999;Krugerand vonSchaewen,2003;

    Ratcliffe and Shachar-Hill, 2006; Schwender et al., 2003),

    there is little information available on the metabolic fluxes of

    secondary metabolism in plants.

    At first glance, it may seem relatively easy to analyse

    linear-type secondary metabolic pathways with limited

    branching. This is true if starting and final rates can be

    measured adequately and if intracellular concentrations are

    known. Unfortunately, this is usually not the case. Precur-

    sors entering the pathway may be produced by metabolic

    pathways at unknown rates, and product accumulation may

    be unable to be measured because appropriate experimen-

    tal techniques are unavailable, or because the products

    undergo further reactions. In such cases, dynamic tech-niques have to be applied. These take into account the

    accumulation of compounds, and are most successfully

    applied if preliminary kinetic information is already avail-

    able. Rohwer andBotha (2001) analysed the accumulation of

    sucrose in sugarcane culms using in vitro kinetic data.

    McNeil et al. (2000) identified key constraints in an engin-

    eered pathway producing glycine betaine. Labelling tech-

    niques combined with simple first-order kinetic models can

    be used to determine metabolic rates, as shown earlier by

    Sims and Folkes (1964). Recently, Matsuda et al. (2003)

    modified this method and applied it to L-phenylalanine (Phe)

    metabolism in potato tubers, to allow determination of flux

    partitioning at the p-coumaroyl CoA branch point. This

    method is particularly attractive because of its simplicity and

    clear definition. It relies, however, on a series of assump-

    tions, square-step input, first-order kinetics and constant

    rates during the labelling experiment, which limits its use.

    Boatrightet al.(2004) studiedin vivobenzenoid metabolism

    in petunia petal tissue using dynamic labelling experiments

    combined with mass spectrometric analysis of labelling.

    They did, however, assume constant rates throughout their

    experiment. Therefore they could not directly extract kinetic

    information using their method. Full kinetic models with

    MichaelisMenten kinetics were applied by Nuccio et al.

    (2000) to determine glycine betaine synthesis in tobacco. In

    their work they used 14C as the labelling isotope. It is

    certainly preferable to use established kinetic equations, but

    these are often not easily available, for example for secon-

    dary metabolite pathways. In such cases it seems best toapply generally applicable kinetic equations such as power-

    law kinetics, which allow many kinetic dependencies

    observed in biological systems, within a certain range, to

    be mimicked (Voit, 2000). The application of dynamic

    models for flux analysis has a significant advantage because

    kinetic models allow straightforward further analysis. It is

    generally possible to obtain hints about expression levels

    and to calculate flux-control coefficients using the models

    (McNeilet al., 2000). Another way of obtaining flux-control

    coefficients is by modifying enzyme activity and measure-

    ment of flux responses, as was done by Ramli et al. (2002)

    using radioactive tracer substances. Basic information on

    the methodology and techniques of network modelling and

    flux analysis can be found in textbooks (e.g. Stephanopou-

    loset al., 1998). A useful detailed overview is provided in a

    recent comprehensive review of metabolic flux analysis in

    plants, in which important technical terms are also

    explained (Ratcliffe and Shachar-Hill, 2006).

    Here we investigate a method for the determination of

    metabolic fluxes, expression levels of enzymes at each

    reaction step, concentrations of unmeasured intermediates,

    and control coefficients of secondary metabolic pathways in

    plants without a priorikinetic knowledge. First-order and

    power-law kinetics are applied to describe changes in

    metabolitesand metabolitelabelling.A morecomplexmodelis reduced to a simpler one, which is used for statistical

    analysis using Monte Carlo simulations and for calculating

    flux-control coefficients. Experimental data used in the

    present study are derived from pulse-labelling experiments

    carried out in our previous study (Matsuda et al., 2005), in

    which incorporation of exogenously applied L-phenylalan-

    ine-d5 into seven phenylpropanoid metabolites in potato

    tubers was determined. In potato (Solanum tuberosum) the

    following metabolites are derived from the phenylpropanoid

    Method for metabolic flux and control analysis 177

    2007 The AuthorsJournal compilation 2007 Blackwell Publishing Ltd, The Plant Journal, (2007), 50, 176187

  • 8/14/2019 Estimation of Metabolic Fluxes

    3/12

    pathway: one hydroxycinnamic acid ester with quinate

    (chlorogenic acid, CGA) and six amides with tyramine,

    octopamine and putrescine: N-p-coumaroyloctopamine

    (p-CO), N-p-coumaroyltyramine (p-CT), N-feruloyloctopam-

    ine (FO), N-feruloyltyramine (FT); caffeoylputrescine (CafP)

    and feruloylputrescine (FP) (Figure 1). p-CO, FO, p-CTandFT

    accumulation is induced by pathogen infection (Clarke, 1982;

    Kelleret al., 1996), wounding of tissues and treatment with

    elicitors (Miyagawaet al., 1998; Negrelet al., 1993; Schmidt

    et al., 1999). Regulation of these pathways is thus of great

    interest, and a better understanding of the pathways could

    lead to the development of plants with improved defence

    responses. Although the metabolic flux of metabolite bio-

    synthesis has been estimated roughly from the amounts of

    metabolites (Schmidt et al., 1998), the activities of biosyn-

    thesis-related enzymes (Matsuda et al., 2000; Negrel et al.,

    1993), the transcription levels of the genes encoding these

    enzymes (Schmidt et al., 1999), and the regulatory mecha-

    nisms of the pathways are poorly understood due to meth-

    odological limitations. Recently, direct determination of the

    metabolic fluxes of this pathway was achieved using pulsefeeding combined with simple metabolic models. Using this

    method, regulation of the phenylpropanoid pathway in

    control plants and those treated with ab-1,3-glucooligosac-

    charide elicitor (laminarin) has been characterized in some

    detail (Matsuda et al., 2003, 2005). However, the metabolic

    model used in that study was too simple to analyse the

    complex metabolic network shown in Figure 1.

    In the present work we develop a method to answer the

    following questions: (i) Is it possible to determine the

    metabolic flux of all reaction steps shown in Figure 1? (ii)

    Are the results obtained earlier by Matsuda et al. (2005) in

    agreement withthe network model results? (iii) Which fluxes

    are modified significantly upon application of the b-1,3-

    glucooligosaccharide elicitor? (iv) Is it possible to generate

    hypotheses about induction repression based on kinetic

    metabolic flux analysis? (v) Is it possible to predict levels of

    unmeasured intermediate metabolites based on simula-

    tions? (vi) What are the rate-controlling reactions in the

    control and elicitor experiments?

    Results and discussion

    Model formulation and simplification

    Model development began from the reaction scheme shown

    in Figure 1. The appropriateness of the pathway was dis-

    cussed in our previous work (Matsudaet al., 2005). In a first

    model with only first-order kinetics, it was found that the

    reactionsfrom phenylalanine to pCACoAcan be lumped into

    one reaction without influencing the results significantly.This is so because the concentrations of intermediates are

    very low, below the detection limit of the HPLCMS method,

    allowing the assumption of steady state for these metabo-

    lites. Additionally, reactions from pCACoA to CGA and those

    from pCACoA to caffeoyl CoA (CafCoA) via p-coumaroyls-

    hikimate and caffeoylshikimate couldbe lumped together for

    the same reason. The uptake of Phe added was described

    betterby a reversible reactionthat reflects a diffusionprocess

    ina simpleway.In additionto thereactionsdepicted inFigure 1,

    COOH

    NH2

    COOH

    NH2

    COCoA

    HO

    COCoA

    HO

    COCoA

    HO

    HO

    H3CO

    HO

    N

    O

    H

    OH

    HO

    N

    O

    H

    OH

    OH

    HO

    N

    O

    H

    OH

    HO

    N

    O

    H

    OH

    OH

    H3CO

    H3CO

    HO

    N NH2

    O

    H

    H3CO

    HO

    N NH2

    O

    H

    HO

    HO

    O

    OHO

    OH

    OH

    HO COOH

    p-coumaric acid

    Cinnamic acid

    L-phenylalanine

    p-coumaroyl CoA

    caffeoyl CoA

    feruloyl CoA

    chlorogenic acid

    p-coumaroyloctopamine

    p-coumaroyltyramine

    feruloyloctopamine

    feruloyltyramine feruloylputrescine

    caffeoylputrescine

    p-coumaroylshikimatep-coumaroylquinate

    caffeoylshikimate

    PAL

    THT

    THT

    C4H

    4CL

    CQT CST

    CST

    CQT

    CCoAOMT

    PHT

    PHT

    C3'H

    kDPhe

    kPAL

    k4CL

    kTHTpCO

    kTHTpCT

    kpCOdeg

    kpCTdeg

    kCST1kCQT1

    kCQT2

    kCST2

    kPHTCafP

    kPHTFP

    kTHTFO

    kTHTFT

    kFOdeg

    kFTdeg kFPdeg

    kCafPdeg

    C3'HkC3'HCGA

    kCCoAOMT

    Phe

    Pheex

    pCACoApCO

    pCT

    FO

    FT

    CafP

    FP

    CGA

    kC4H

    CafCoA

    FCoA

    kCGAdeg

    kC3'HCafS

    Figure 1. Reaction schemeof themetabolic path-

    ways in potato that start from L-phenylalanine

    (Phe).

    Enzymes are abbreviated as follows: 4CL, 4-hydroxy-

    cinnamic acid:CoA ligase; Phe, phenylalanine ammo-

    nia lyase; PHT, putrescine hydroxycinnamoyl

    transferase;CQT, quinatehydroxycinnamoyl-CoA:qui-

    nate hydroxycinnamoyltransferase; THT, tyramine

    hydroxycinnamoyl-CoA:tyramine hydroxycinnamoyl-

    transferase; C3H, 5-O-(4-hydroxycinnamoyl)quinate3-hydroxylase; C4H, cinnamic acid 4-hydroxylase;

    CST, shikimate hydroxycinnamoyl-CoA:shikimate

    hydroxycinnamoyltransferase.

    178 E. Heinzleet al.

    2007 The AuthorsJournal compilation 2007 Blackwell Publishing Ltd, The Plant Journal, (2007), 50, 176187

  • 8/14/2019 Estimation of Metabolic Fluxes

    4/12

    three hypothetical reactions from hydroxycinnamoyl CoAs

    (pCACoA, CafCoA and FCoA) into other metabolites or poly-

    mers were introduced as indicated by their corresponding

    rateconstants, kpCACoAdeg, kCafCoAdeg and kFCoAdeg (Figure 2a),

    as the hydroxycinnamoyl CoAs could serve as precursors forthe synthesis of other defence-related polymers, including

    lignin, which might be upregulated on elicitor treatment.

    From this, the so called first-order model was obtained (Fig-

    ure 2a). Using this model and the experimental data from

    Matsudaet al.(2005), we found that the reactions indicated

    by the rate constants kpCACoAdeg, kCafCoAdeg and kFCoAdeg were

    always negligibly small. These were therefore eliminated

    from further studies, resulting in the scheme shown in

    Figure 2(b), which we named the power-law model.

    In the next step, all parameters (11 initial concentrations,

    20 rate constants and 11 exponents of power-law kinetics)

    were estimated using weighted least squares. Initial values

    were estimated for all concentrations, also for those that

    were experimentally observed, because they also had an

    experimental error of about 20%. Details of weighting are

    given in Experimental procedures. The whole calculation

    procedure is summarized schematically in Figure 3.

    Using Berkeley Madonna, these parameters were fitted to

    the 68 experimental points of each experiment for the

    control and elicitor treatments (Tables S1S3). It was poss-

    ible to fit the experimental data very well, as can be seen in

    Figure 4 for the labelling data.

    Each estimation required about 500010 000 runs, which

    were finished in

  • 8/14/2019 Estimation of Metabolic Fluxes

    5/12

    including JPAL, JCQT2, JCGAdeg, JCST1 and JCCoAMT, which

    could not be determined in our previous study due to excess

    simplification of the metabolic model. This improvement is

    an obvious advantage of the present method, by which a

    more detailed flux distribution in the metabolic network was

    determined. Moreover, the estimated values of metabolic

    fluxes showed good agreement with those determined in

    our previous study (Table 1; Matsuda et al., 2005), except for

    the degradation steps includingJpCOdeg,JpCTdeg,JpFOdegand

    JpFTdeg. One possible reason for the discrepancy is that

    degradation reactions are generally predicted with higher

    uncertainty (Table 1; Figure 5).

    In the control case, 74% of Phe is converted to pCACoA,

    and most of it ends up in FO (45%), although 19% is

    converted to FP, 10% to CafP and about 4% to FT. Only 16%

    of Phe is converted to pCO, and 6% to pCT. Only FO and FP

    seem to be converted to further species, but the estimations

    of these fluxes are characterized by high standard devia-

    tions. In the potato tuber disks to which elicitor was applied,

    the fluxes JPAL, JTHTpCO, JTHTpCT and JCGAdeg increased

    significantly. The mean values of fluxes from pCO and pCT

    to other products, JpCOdeg and JpCTdeg, also increased

    significantly. Other reaction rates, JCQT1, JCST1, JCCoAOMT,JPHTCafP, JTHTFO and JPHTFO, clearly decreased. This means

    that, after elicitor treatment, the metabolism is shifted to

    pCO and pCT biosynthesis, compounds that are further

    converted to polymer compounds to form a physical barrier

    against pathogen intrusion (Schmidt et al., 1998). These

    results also suggest that the role of pCO in defence response

    is more important than that of other compounds, which

    generally agrees with the results of our previous study

    (Matsudaet al., 2005).

    Kinetic analysis and expression

    A second set of answers obtained from the model is based

    on its kinetic properties. Figure 6 shows the rate constants.

    These are the product of enzyme concentration and the

    actual rate constant, and might also be influenced generally

    Figure 3. Procedures for modelling and param-

    eter estimation.Q is the weighted least squares,

    yithe experimental value, Pthe parameter vec-

    tor,xithe independent variable,f(xi,P) are model

    values, andwi= 1/ri is the weighting factor.

    Phe pCO pCT CGA

    0 5

    CafP

    0 5

    FO

    0 5

    FT

    0 5

    FP

    Phe pCO pCT CGA

    0 5

    CafP

    0 5

    FO

    0 5

    FT

    0 5

    FP

    Time (h)

    Isotopeabundance

    Isotopeabundance

    Time (h)

    (b)

    (a)

    0

    0.2

    0.4

    0.6

    0.8

    1

    0

    0.2

    0.4

    0.6

    0.8

    0

    0.2

    0.4

    0.6

    0.8

    0

    0.2

    0.4

    0.6

    0.8

    Figure 4. Fractional 13C enrichment calculated by the model compared with

    experimental data (symbols) using the data set obtained from control (a) and

    elicitor-treated (b) samples.

    180 E. Heinzleet al.

    2007 The AuthorsJournal compilation 2007 Blackwell Publishing Ltd, The Plant Journal, (2007), 50, 176187

  • 8/14/2019 Estimation of Metabolic Fluxes

    6/12

    by other metabolites. First, it can be seen that these con-

    stants are determined with relatively small errors, and vary

    over orders of magnitude. Assuming that the effects of other

    metabolites are negligible, predictions of expression-

    induction phenomena can be made directly. This can be

    done at each metabolic branch point, for example at pCA-

    CoA, CafCoA and FCoA (Figure 7), by observing the relative

    changes. Even if the estimation of the concentration of each

    starting metabolite was not very precise, the relative

    magnitudes of the rate constants clearly indicate induction-

    repression phenomena. These can be seen in comparative

    linear plots ofkTHTpCO, kTHTCpT, kCQT1 and kCST1 in Figure 7(a).

    Both constants kTHTpCO and kTHTpCT increased significantly,

    18- and 10-fold, respectively, whereas bothkCQT1and kCST1remained constant. The activation of THT activity in b-1,3-

    gluco-oligosaccharide elicitor-treated potato tubers has

    been observed earlier in independent experiments (Matsuda

    et al., 2000), and thus supports our findings. At the same

    time, the corresponding fluxes JTHTpCO and JTHTpCT also in-

    creased, whereas JCQT1 and JCST1 both decreased signifi-

    cantly, as shown in Figure 5. This can be explained by the

    estimated reduced concentration of pCACoA, as shown in

    Figure 8. The branch point at CafCoA showed a significant

    increase in kCCoAOMT, whereas kPHTCafP remained nearly

    constant (Figure 7b). Both associated fluxes, JPHT1CafP and

    JCCoAOMT, decreased upon elicitation, as shown in Figure 5.

    The third branch point studied, the reactions starting from

    FCoA, showed increases ofkTHTFO and kTHTFT, which were,

    however, not significant considering the large degree of

    error, and a significant decrease ofkPHTFP as shown in Fig-

    ure 7(c). The fluxes JTHTFO and JPHTFP decreased, whereasJTHTFT remained about constant (Figure 5). These results

    clearly show the regulation effects in thismetabolic network.

    Flux-control analysis

    The model was also used further for flux-control analysis

    (Table 2). The concept of flux-control coefficients (Fell, 1997;

    Kacser and Burns, 1973) was introduced to estimate

    quantitatively the contribution of each metabolic step to

    Table 1 Estimated absolute and relative metabolic fluxes, Ji, and their standard deviations for the control and for the case with elicitor

    application

    Control Elicitor

    Mean flux

    [nmol (g FW)1 h1]

    Relative flux

    (%)

    Flux determined

    by Matsuda et al.

    (2005) [nmol (g FW)1 h1]

    Mean flux

    [nmol (g FW)-1 h1]

    Relative flux

    (%)

    Flux determined

    by Matsuda et al.

    (2005) [nmol (g FW)1 h1]

    JPAL 8.59 0.41 100 0.0 nda 11.87 0.89 100 0.0 nd

    JTHTpCO 1.41 0.22 16.4 2.8 0.98 7.6 0.6 64.0 0.3 8.7

    JpCOdeg 0.00 0.00 0.0 0.0 0.98 2.69 1.73 21.2 17.2 7.4

    JTHTpCT 0.75 0.24 8.6 2.7 0.34 2.14 0.15 18.0 0.3 2.5

    JpCTdeg 0.00 0.00 0.0 0.0 0.33 1.07 0.37 9.1 3.7 2.3

    JCQT1 0.34 0.03 4.0 0.3 0.38 0.12 0.01 1.0 0.0 0.14

    JCGAdeg 0.03 0.00 0.3 0.0 nd 0.16 0.01 1.4 0.1 nd

    JCQT2 0.23 0.33 2.6 3.8 nd 0.47 0.2 4.1 1.9 nd

    JCST1 6.10 0.41 71 3.2 nd 2.01 0.16 17.0 0.4 nd

    JCCoAOMT 5.54 0.55 64.4 5.6 nd 2.42 0.2 20.7 1.8 nd

    JPHTCafP 0.79 0.08 9.2 0.8 1.14 0.05 0.01 0.4 0.0 0.14

    JCafPdeg 0.12 0.20 1.4 2.5 0.49 0.55 0.27 4.5 2.6 0.11

    JPHTFP 1.63 0.19 19.0 2.3 1.22 0.11 0.02 0.9 0.2 0.14

    JFPdeg 0.04 0.07 0.5 0.8 0.13 0.12 0.19 1.2 1.8 0.04

    JTHTFO 3.59 0.41 41.7 4.1 3.4 2.05 0.19 17.4 1.4 3.4

    JFOdeg 0.42 0.83 4.8 9.4 3.3 0.79 0.68 7.4 6.5 3.3

    JTHTFT 0.33 0.08 3.9 0.9 0.26 0.26 0.04 2.3 0.5 0.59

    JFTdeg 0.00 0.00 0.0 0.0 0.24 0.01 0.02 0.1 0.2 0.52

    Metabolic flux values determined from the same data set using a different method by Matsuda et al.(2005). a, not determined.

    Phe

    pCACoA

    CGA

    CafCoA

    FCoA

    CafP

    FP

    pCOpCT

    JCST1

    JPAL

    rCQT1

    JCQT2

    JpCOdeg

    JpCTdeg

    JTHTpCO

    JTHTpCT

    FOFT

    JFOdeg

    JFTdeg

    JTHTFO

    JTHTFT

    JCGAdeg

    JPHTcafP

    JPHTFP

    JcafPdeg

    JFPdeg

    JCCoAOMT3.6 0.4/ 2.0 0.20.4 0.8/0.8 0.7

    0.0/0.01 0.02 0.1/0.3 0.00.3

    1.6 0.2/0.1 0.0

    5.5 0.6/2.4 0.2

    0.04 0.07/0.1 0.2

    0.1 0.2/ 0.6 0.30.8 0.1/ 0.1 0.0

    6.1 0.4 / 2.0 0.2

    0.2 1.7/ 0.5 0.2

    0.03 0.00/0.2 0.0

    0.3 0.0/0.1 0.0

    0.8 0.2/ 2.1 0.2

    1.4 0.2/ 7.6 0.60.0 / 2.7 1.7

    0.0 / 1.1 0.4

    8.6 0.4 / 11.9 0.9

    Control/elicitor

    (nmol(g FW)1 h1)

    Figure 5. Absolute fluxes for the control and elicitor cases. Values for the

    control and for the case with elicitor application are shown at the corres-

    ponding reaction steps (control/elicitor), in nmol (g FW)1 h1.

    Method for metabolic flux and control analysis 181

    2007 The AuthorsJournal compilation 2007 Blackwell Publishing Ltd, The Plant Journal, (2007), 50, 176187

  • 8/14/2019 Estimation of Metabolic Fluxes

    7/12

    controlling a metabolic flux in a pathway. For JTHTpCOof the

    control sample (Table 2), control coefficients of the steps for

    kDPheandkTHTpCOhad large positive values. This means that

    the increase inkTHTpCO(in other words, the activation of the

    enzyme responsible for this step) is mainly responsible for

    the increase in JTHTpCO. In contrast, control coefficients for

    kTHTpCT, kCQT1 and kCST1 were negative, indicating that an

    increase in their values caused a decrease in JTHTpCO.

    However, the step ofkCST1is more important thankTHTpCTand kCQT1 for controlling the metabolic flux ofJTHTpCO, asthe

    control coefficient ofkCST1()73) was much smaller than the

    control coefficients of kTHTpCT and kCQT1 ()13 and )5,

    respectively). In Table 2, the sum of the values of the controlcoefficients in each row is always 100. It is obvious that the

    supply ofL-phenylalanine was important for most reactions

    of the system, as it is the only input. Only for minor

    metabolic activities where accumulation played a dominant

    role (e.g. JCGAdeg, JCQT2, JCafPdeg, JFPdeg, JFOdegand JFTdeg)

    was it less important. Most interesting here is a comparison

    of control and elicitor cases to determine whether flux

    control has moved to other reactions after elicitor applica-

    tion. JTHTpCO was mostly controlled by itself and was in

    competition withJCST1. The competition was smaller for the

    elicitor case. JTHTpCT was mostly in competition with the

    reaction via p-coumaroylshikimate and, after elicitor appli-

    cation, with the reaction to pCO. Significant changes were

    observed for JCQT1. Competition from the reaction to pCO,

    JTHTpCO, increased and competition from JCST1 decreased.

    JCQT2was influenced primarily by kCQT2. The flux JCST1was

    increasingly in competition with JTHTpCO after elicitor

    Figure 6. Estimated kinetic constants of the power-law model (Figure 2) with

    fixed values for exponents of power-law kinetics as specified in the text. The k

    values represent the rate constants of each reaction step as defined by Eqn 3

    in Experimental procedures. Error bars, standard deviations derived from

    Monte Carlo simulations.

    0

    5

    10

    15

    20

    25

    30

    kTHTFO kTHTFT kPHTFP

    Control

    Elicitor

    0

    50

    100

    150

    200

    250

    300

    350

    400

    450

    500

    kTHTpCO kTHTpCT kCQT1 kCST1

    Control

    Elicitor

    0

    200

    400

    600

    800

    1000

    1200

    1400

    kPHTCafP kCCoAOMT

    Rateconstant(h1)

    Rateconsta

    nt(h1)

    Rateconstant(h

    1)

    Control

    Elicitor

    (a)

    (b)

    (c)

    Figure 7. Comparison of kinetic constants around various nodes.

    (a) p-coumaroyl CoA (pCCoA); (b) caffeoyl CoA (CafCoA); (c) feruloyl CoA

    (FCoA). Kinetic constants are defined in Figure 2. Error bars, standard

    deviations derived from Monte Carlo simulations.

    182 E. Heinzleet al.

    2007 The AuthorsJournal compilation 2007 Blackwell Publishing Ltd, The Plant Journal, (2007), 50, 176187

  • 8/14/2019 Estimation of Metabolic Fluxes

    8/12

    application. The reaction to CafP, JPHTCafP, was primarily in

    competition with the reaction to FCoA, to a slightly lesser

    extent with the reaction to pCO, and to an even lesser extent

    with the reaction to pCT; however, it was stimulated by

    JCST1, especially after elicitation, when kCQT2 also had a

    stimulating effect. The reactions to THTFO and THTFT were

    similarly influenced by all preceding reactions, but were

    main competitors with each other. To a lesser extent, JPHT2

    was also a competitor. The degradation reactions weregenerally predicted with higher uncertainty (Table 1;

    Figure 5). Therefore information about their control is also

    generally less accurate. Their control coefficients did not

    change very much upon elicitation.

    The results of the metabolic control analysis revealed

    potential targets for rational metabolic engineering. As

    elicitation simulates the plants defences against pathogen

    infection (Clarke, 1982; Keller et al., 1996) and wounding of

    tissues (Miyagawaet al., 1998; Negrel et al., 1993; Schmidt

    et al., 1999), amplification ofJpCOdegis desirable. This could

    be accomplished by amplification ofp-coumaroyloctopam-

    ine-synthesizing and -degrading enzymes, characterized by

    kTHTpCOandkpCOdeg, which both have a highly positive flux-

    control coefficient for JpCOdeg. A reduction of the flux

    towardsp-coumaroyltyramine would also be beneficial.

    Conclusions

    We have shown in the present study that flux analysis using

    labelled precursors is possible, even under dynamic condi-

    tions, withouta prioriknowledge of kinetics. This is possible

    by using power-law kinetics, which can describe almost any

    kinetics with sufficient accuracy. The analysis completely

    answers the questions introduced in the Introduction.

    Dynamic tracer experiments not only allow the estimation of

    metabolic fluxrates (Table 1) and the distribution of network

    activities (Figure 5), but also permit further direct analysis,

    for example of flux-control coefficients (Table 2). Flux-con-

    trol coefficients can be used to identify targets for engin-

    eering plant metabolism. In the present study, suggestions

    for improved plant self-protection could be elaborated. The

    application of dynamic models in combination with statis-

    tical model evaluation allows additional interpretation, for

    example, prediction of unmeasured intermediate concen-

    trations (Figure 8) and prediction of expression/repression

    phenomena (Figures 6 and 7). Such dynamic models can be

    applied even for rather complex systems using the fast and

    reliable computational tools available.

    We believe this methodology will also be very useful for

    large-scale quantitative analysis of metabolic networks after

    the application of pulses or step changes of labelled water,carbon dioxide or ammonia (e.g. 2H2O, H2

    18O, 13CO2 or15NH3). Then it would be possible to analyse simultaneously

    a whole set of sub-networks using the methodology pre-

    sented here, provided corresponding metabolite analysis is

    available (Soga et al., 2002). The method of metabolic

    analysis demonstrated in this study is also expected to be

    a powerful tool for investigation of plant metabolic func-

    tions, as well as for rational metabolic engineering.

    Experimental procedures

    Model formulation

    The pathway investigated here is depicted in Figure 1, and is des-

    cribed elsewhere in greater detail (Matsuda et al., 2003, 2005).

    Starting from this reaction network, the model equations were set

    up using standard techniques described in textbooks (e.g. Dunn

    et al., 2003). In all formulations, a well-mixed system is assumed.

    Material balances were formulated for the constant volume system

    for each labelled and non-labelled species:

    dCj;k

    dt Xn

    j1

    Jj;k 1

    where Cis concentration, i indicates the component, k indicates

    whether the component is a labelled or non-labelled species, Ji s

    metabolic flux or reaction rate, and j indicates reactions producingand consuming component i.

    The balance for labelled pCACoA of the power-law model

    (Figure 2b) is, for example:

    dCpCACoA;i

    dt JPAL;i JTHTpCO;i JTHTpCT;i JCQT1;i JCST1;i 2

    whereJrepresents the reaction rates of fluxes each formulated as:

    Jj kjCaji k

    0jC

    ajE;jC

    aji 3

    0

    0.02

    0.04

    0.06

    0.08

    0.1

    0.12

    0.14

    0.16

    pCACoA CafCoA FCoa

    Concentration(nmol(gFW)1)

    Control

    Elicitor

    Figure 8. Estimated average concentrations of unmeasured compounds p-coumaroyl CoA(pCCoA),caffeoyl CoA(CafCoA)and feruloyl CoA(FCoA), with

    error bars derived from Monte Carlo simulations.

    Method for metabolic flux and control analysis 183

    2007 The AuthorsJournal compilation 2007 Blackwell Publishing Ltd, The Plant Journal, (2007), 50, 176187

  • 8/14/2019 Estimation of Metabolic Fluxes

    9/12

    Here jindicates the reaction, i the reacting component and E the

    enzyme. For the reactions indicated by thin lines in Figure 2(b), only

    first-order kinetics were used (ai= 1). For the reactions indicated by

    thick lines in Figure 2(b), power-law kinetics were applied. In this

    simplified formulation, it is assumed that influences from other

    substances, either co-substrates or other compounds allosterically

    Table 2Flux-control coefficients estimated using the model

    KDPhe kPAL kTHTpCO kpCOdeg kTHTpCT kpCTdeg kCQT1 kCGAdeg kCQT2 kCST1 kCCoAOMT kPHTCafP kCafPdeg kPHTFP kFPdeg kTHTFO kFOdeg kTHTFT kFTdeg

    JPAL

    C 97 3 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

    E 100 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

    JTHTpCO

    C 105 4 84 0 )

    13 0 )

    5 0 0 )

    74 0 0 0 0 0 0 0 0 0E 101 0 35 0 )19 0 )1 0 0 )17 0 0 0 0 0 0 0 0 0

    JpCOdeg

    C 84 5 69 96 )11 0 )4 0 0 )60 0 0 0 0 0 0 0 0 0

    E 51 1 18 79 )10 0 )1 0 0 )9 0 0 0 0 0 0 0 0 0

    JTHTpCT

    C 105 4 )16 0 87 0 )5 0 0 )74 0 0 0 0 0 0 0 0 0

    E 101 0 )65 0 81 0 )1 0 0 )17 0 0 0 0 0 0 0 0 0

    JpCTdeg

    C 96 5 )15 0 81 100 )5 0 0 )69 0 0 0 0 0 0 0 0 0

    E 65 1 )44 0 55 57 )1 0 0 )12 0 0 0 0 0 0 0 0 0

    JCQT1

    C 116 4 )18 0 )15 0 95 0 0 )82 0 0 0 0 0 0 0 0 0

    E 113 0 )72 0 )21 0 99 0 0 )19 0 0 0 0 0 0 0 0 0

    JCGAdeg

    C 8 0 )

    1 0 )

    1 0 7 100 0 )

    6 0 0 0 0 0 0 0 0 0E 5 0 )3 0 )1 0 5 93 )13 )1 0 0 0 0 0 0 0 0 0

    JCQT2

    C 7 0 )1 0 )1 0 6 0 100 )5 0 0 0 0 0 0 0 0 0

    E 5 0 )3 0 )1 0 4 )6 88 )1 0 0 0 0 0 0 0 0 0

    JCST1

    C 93 3 )14 0 )12 0 )4 0 0 34 0 0 0 0 0 0 0 0 0

    E 90 0 )58 0 )17 0 )1 0 0 85 0 0 0 0 0 0 0 0 0

    JCCoAOMT

    C 91 3 )14 0 )11 0 )4 0 0 33 15 )12 0 0 0 0 0 0 0

    E 77 0 )49 0 )14 0 0 )1 13 72 2 )2 0 0 0 0 0 0 0

    JPHTCafP

    C 109 4 )16 0 )14 0 )5 0 0 40 )102 85 0 0 0 0 0 0 0

    E 92 0 )59 0 )17 0 0 )1 16 86 )117 98 0 0 0 0 0 0 0

    JCafPdeg

    C 22 1 )

    3 0 )

    3 0 )

    1 0 0 8 )

    21 18 86 0 0 0 0 0 0E 2 0 )2 0 0 0 0 0 0 2 )3 2 66 0 0 0 0 0 0

    JPHTFP

    C 106 5 )16 0 )13 0 )5 0 0 40 18 )15 0 62 0 )73 0 )8 0

    E 97 0 )62 0 )18 0 0 )1 17 90 3 )3 0 94 0 )105 0 )15 0

    JFPdeg

    C 28 2 )4 0 )3 0 )1 0 0 10 5 )4 0 17 99 )19 0 )2 0

    E 4 0 )3 0 )1 0 0 0 1 4 0 0 0 4 98 )5 0 )1 0

    JTHTFO

    C 79 4 )12 0 )10 0 )4 0 0 30 13 )11 0 )29 0 46 0 )6 0

    E 73 0 )47 0 )13 0 0 )1 13 68 2 )2 0 )5 0 22 0 )12 0

    JFOdeg

    C 67 4 )10 0 )8 0 )3 0 0 25 11 )9 0 )24 0 39 99 )5 0

    E 32 1 )21 0 )6 0 0 0 6 31 1 )1 0 )2 0 10 74 )5 0

    JTHTFT

    C 106 5 )

    16 0 )

    13 0 )

    5 0 0 40 18 )

    15 0 )

    38 0 )

    73 0 92 0E 97 0 )62 0 )18 0 0 )1 17 90 3 )3 0 )6 0 )105 0 85 0

    JFTdeg

    C 60 3 )9 0 )8 0 )3 0 0 23 10 )8 0 )22 0 )41 0 53 100

    E 33 1 )22 0 )6 0 0 0 6 32 1 )1 0 )2 0 )37 0 30 92

    Control coefficients are given as percentages. C, control experiment; E, elicitor treatment.

    184 E. Heinzleet al.

    2007 The AuthorsJournal compilation 2007 Blackwell Publishing Ltd, The Plant Journal, (2007), 50, 176187

  • 8/14/2019 Estimation of Metabolic Fluxes

    10/12

    influencing rates, are constant throughout one experiment. These

    influences would then be summarized in the value ofkj. The rates of

    the labelled and non-labelled compounds were calculated by:

    Jj;kJjCi;k

    Ci4

    with variables as defined above.

    Flux experiments

    All experiments were carried out by Matsuda et al. (2003, 2005)

    and are described in detail there. Here we describe only essentials

    for setting up the model. Disk-shaped slices of potato tubers were

    incubated in the dark at 18C. After 24 h the elicitor laminarin was

    applied to a subset of disks. After a further 12 h, labelled L-phe-

    nyl-d5-alanine was added to all disks. For concentration and

    labelling analysis, three potato tuber disks were analysed by li-

    quid chromatographymass spectroscopy (LCMS). Labelling data

    were used directly for parameter estimation. Additional experi-

    mental data used were initial concentrations (concentrations at

    the time when labelling was applied, 36 h after preparation of

    disks). To take into consideration the past time-course of the

    concentration changes, initial rates of concentration changes

    were calculated from the slopes of concentration at time zero in

    the labelling experiment (36 h after disk preparation and 12 h

    after elicitor application). Experimental data are given in

    Tables S1S3.

    Simulation and parameter estimation

    The calculation procedures are summarized schematically in

    Figure 3. Initial simulations were carried out using Berkeley Ma-

    donna (http://www.berkeleymadonna.com; Dunn et al., 2003).

    This software is fast and allowed quick parameter estimation

    using available integrators for stiff systems. Rate constants and

    initial concentrations were constrained to values 0, and expo-

    nents of power-law equations were constrained to 0.2 < a < 1.5.

    Weighting was carried out using experimentally identified stan-dard deviations. These were: concentrations, 20%; initial rates,

    10%; fractional labelling, 10%. For initial rates the error was

    calculated as a percentage of the initial concentrations. Data used

    are specified in Table S1. It was found that parameter estimation

    depended quite strongly on the initial conditions. Calculations

    with initial conditions far from the final values did not always

    converge to the global minimum. This lack of convergence could,

    however, be seen easily from graphical comparison of simulation

    results and experimental data, and also from the calculated

    weighted sum of square deviations. Under conditions where all

    experimental data fitted reasonably well, the identified minima

    were very close.

    In the first set of simulations, all initial concentrations, initial

    rates and kinetic constants could be fitted simultaneously. This

    was necessary because initial conditions and initial rates were

    also corrupted by experimental errors and had to be estimated.

    Convergence was usually obtained after 5000 to 10 000 simula-

    tions. In later simulations, initial conditions and exponents of the

    power-law kinetics were fixed and only rate constants were

    estimated. The results found were confirmed by simulation using

    MATLAB. MATLAB simulations used the ode15 s integrator and

    fminsearch (unconstrained optimization) or fmincon (constrained

    optimization) as optimizers. The MATLAB simulations were con-

    siderably slower than those carried out using Berkeley Madonna.

    Also, in the MATLAB optimizations convergence depended on the

    choice of initial conditions: having initial conditions sufficiently

    close to the final values always gave very similar results. MATLAB

    was also used for further analysis of parameter estimation

    results. These included the sensitivities of estimation for param-

    eter values, as well as pairwise contour plots of the sum of

    weighted square deviations as a function of two parameters to

    check for parameter correlation. Final statistical analysis of the

    model and results was carried out using Monte Carlo simula-

    tions. All measurement data were varied according to theirindividual error ranges using the nrmrnd function of MATLAB. In

    these runs, exponents of power-law kinetics were fixed and only

    rate constants were estimated. Starting values for estimated

    parameters and initial values were those identified in Berkeley

    Madonna simulations.

    Flux-control analysis

    Flux-control coefficients were originally introduced by Kacser and

    Burns (1973) and later described in more detail by Fell (1997):

    CJkj @Jk

    @Ej

    Ej

    Jk

    @lnJk

    @lnEj5

    Fluxes as defined in Eqn 3 are identical to reaction rates. Assuming

    that the rate constants k defined in Eqn 3 are proportional to en-

    zyme concentration, we obtain:

    CJkj @Jk

    @Ej

    Ej

    Jk

    @Jk

    @Kj

    Kj

    Jk6

    Thederivative Jk/kj was calculated numerically by calculatingfluxes at a total of nine equally spaced parameter values 20%around the optimal value. After smoothing using the csapsspline

    tool of MATLAB, the derivative was calculated at the optimal

    parameter value and was eventually estimated as a percentage. The

    whole MATLABcode necessary to calculate flux-control coefficients

    is contained in the Supplementary material. This also includes the

    kinetic model, as well as all data used for the calculations.

    Acknowledgements

    The first author would like to express his deep gratitude for the

    generous offer of being a guest professor in the Agricultural

    Department of Kyoto University. This work was supported in part by

    a grant from the Ministry of Education, Culture, Sports, Science, and

    Technology of Japan (T.N.),and by the CREST programof the Japan

    Science and Technology Organization.

    Supplementary material

    The following supplementary material is available for this article

    online:Table S1 Label data for metabolites determined by LCMS taken

    from Matsudaet al. (2005).

    Table S2 Label time-course forphenylalanine determined by LCMS

    (Matsudaet al., 2005).

    Table S3 Initial concentration and initial rates taken from Matsuda

    et al.(2005).

    Appendix S1 Set ofMATLABm-files for the calculation of flux-control

    coefficients.

    This material is available as part of the online article from http://

    www.blackwell-synergy.com.

    Method for metabolic flux and control analysis 185

    2007 The AuthorsJournal compilation 2007 Blackwell Publishing Ltd, The Plant Journal, (2007), 50, 176187

  • 8/14/2019 Estimation of Metabolic Fluxes

    11/12

    References

    Bailey, J. E. (1998) Mathematical modeling and analysis in bio-

    chemical engineering: past accomplishments and future oppor-

    tunities. Biotechnol. Prog.14, 820.

    Boatright, J., Negre, F., Chen, X., Kish, C. M., Wood, B., Peel, G.,

    Orlova, I., Gang, D., Rhodes, D. and Dudareva, N. (2004) Under-

    standing in vivobenzenoid metabolism in petunia petal tissue.

    Plant Physiol. 135, 19932011.Clarke, D. D. (1982) The accumulation of cinnamic acidamidesin the

    cell walls of potato tissue as an early response to fungal attack. In

    Active Defence Mechanisms in Plants(R. Wood, ed.). New York:

    Plenum Press, pp. 321322.

    Dauner, M. and Sauer, U. (2000) GCMS analysis of amino acids

    rapidly provides rich information for isotopomer balancing.Bio-

    technol. Prog.16, 642649.

    Dunn,I. J.,Heinzle,E., Ingham, J. andPrenosil, J. E. (2003)Biological

    Reaction Engineering. Dynamic Modelling Fundamentals with

    Simulation Exercises, 2nd edn. Weinheim, Germany: Wiley-VCH.

    Fell, D. (1997) Understanding the Control of Metabolism. London:

    Portland Press.

    Giersch, C. (2000) Mathematical modelling of metabolism. Curr.

    Opin. Plant Biol.3, 249253.

    Glawischnig, E., Gierl, A., Tomas, A., Bacher, A. and Eisenreich, W.(2002) Starch biosynthesis and intermediary metabolism in maize

    kernels. Quantitative analysis of metabolite flux by nuclear

    magnetic resonance.Plant Physiol.130, 17171727.

    Hanson, A. D. and Shanks, J. V. (2002) Plant metabolic engineering.

    Entering the S curve.Metab. Eng.4, 12.

    Hellerstein, M. K. (2004) New stable isotopemass spectrometric

    techniques for measuring fluxes through intact metabolic path-

    ways in mammalian systems: introduction of moving pictures

    into functional genomics and biochemical phenotyping. Metab.

    Eng. 6, 85100.

    Hellerstein, M. K. and Neese, R. A. (1999) Mass isotopomer distri-

    bution analysis at 80 years: theoretical, analytic, and experi-

    mental considerations.Am. J. Physiol.276, E1146E1170.

    Kacser, H. and Burns, J. (1973) The control of flux. Symp. Soc. Exp.

    Biol. 27, 65104.Keller, H., Hohlfeld, H., Wray, V., Hahlbrock, K., Scheel, D. and

    Strack, D.(1996) Changes in the accumulation of soluble and cell

    wall-bound phenolics in elicitor-treated cell suspension cultures

    and fungus-infected leaves of Solanum tuberosum. Phytochem-

    istry,42, 389396.

    Kruger, N. J. and von Schaewen, A. (2003) The oxidative pentose

    phosphate pathway: structure and organisation.Curr. Opin. Plant

    Biol. 6, 236246.

    Matsuda, F., Miyagawa, H. and Ueno, T. (2000) Beta-1,3-glucoo-

    ligosaccharide induced activation of four enzymes responsible

    for N-p-coumaroyloctopamine biosynthesis in potato (Solanum

    tuberosum cv.) tuber tissue. Z. Naturforsch. C,55, 373382.

    Matsuda, F., Morino, K., Miyashita, M. and Miyagawa, H. (2003)

    Metabolic flux analysis of the phenylpropanoid pathway in

    wound-healing potato tuber tissue using stable isotope-labeledtracer and LCMS spectroscopy. Plant Cell Physiol.44, 510517.

    Matsuda, F., Morino, K., Ano, R., Kuzawa, M., Wakasa, K. and

    Miyagawa, H. (2005) Metabolic flux analysis of the phenylprop-

    anoid pathway in elicitor-treated potato tuber tissue. Plant Cell

    Physiol. 46, 454466.

    McNeil, S. D., Rhodes, D., Russell, B. L., Nuccio, M. L., Shachar-Hill,

    Y. and Hanson, A. D. (2000) Metabolic modeling identifies key

    constraints on an engineered glycine betaine synthesis pathway

    in tobacco.Plant Physiol. 124, 153162.

    Miyagawa, H., Ishihara, A., Lim, C.-H., Ueno, T. and Furuichi, N.

    (1998) Induction of N-p-coumaroyloctopamine in potato tuber

    disks byb-1,3-glucooligosaccharide. J. Pestic. Sci. 23, 4953.

    Munack, A. (1987) Optimal strategy for identification of monod-type

    models by fed-batch experiments. Computer application in fer-

    mentation technology. InModelling and Control of Biochemical

    Processes (N. Fish, R. Fox and N. Thornhill, eds). Barking, UK:

    Elsevier, pp. 195204.

    Negrel, J., Javelle, F. and Paynot, M. (1993) Wound-inducedtyramine hydroxycinnamoyl transferase in potato (Solanum

    tuberosum) tuber discs. J. Plant Physiol.142, 518524.

    Nielsen, J.(1998) Metabolic engineering: techniques for analysis of

    targets for genetic manipulations. Biotechnol. Bioeng. 58, 125

    132.

    Nuccio, M. L., McNeil, S. D., Ziemak, M. J., Hanson, A. D., Jain, R. K.

    and Selvaraj, G. (2000) Choline import into chloroplasts limits

    glycine betaine synthesis in tobacco: analysis of plants engin-

    eered with a chloroplastic or a cytosolic pathway.Metab. Eng.2,

    300311.

    Ramli, U. S., Baker, D. S., Quant, P. A. and Harwood, J. L. (2002)

    Control analysis of lipid biosynthesis in tissue cultures from oil

    crops shows that flux control is shared between fatty acid syn-

    thesis and lipid assembly.Biochem. J.364, 393401.

    Ratcliffe, R. G. and Shachar-Hill, Y. (2006) Measuring multiple fluxesthrough plant metabolic networks. Plant J.45, 490511.

    Rohwer, J. M. and Botha, F. C. (2001) Analysis of sucrose accumu-

    lation in the sugar cane culm on the basis ofin vitrokinetic data.

    Biochem. J.358, 437445.

    Rontein, D., Dieuaide-Noubhani, M., Dufourc, E. J., Raymond, P.

    and Rolin, D. (2002) The metabolic architecture of plant cells.

    Stability of central metabolism and flexibility of anabolic path-

    ways during the growth cycle of tomato cells.J. Biol. Chem.277,

    4394843960.

    Roscher, A., Kruger, N. J. and Ratcliffe, R. G. (2000) Strategies for

    metabolic flux analysis in plants using isotope labelling. J. Bio-

    technol. 77, 81102.

    Schmidt, A., Scheel, D. and Strack, D. (1998) Elicitor-stimulated

    biosynthesis of hydroxycinnamoyltyramines in cell suspension

    cultures ofSolanum tuberosum.Planta,205, 5155.Schmidt, A., Grimm, R., Schmidt, J., Scheel, D., Strack, D. and

    Rosahl, S. (1999) Cloning and expression of a potato cDNA

    encoding hydroxycinnamoyl-CoA:tyramine N-(hydrox-

    ycinnamoyl)transferase. J. Biol. Chem. 274, 42734280.

    Schwender, J., Ohlrogge, J. B. and Shachar-Hill, Y. (2003) A flux

    model of glycolysis and the oxidative pentosephosphatepathway

    in developing Brassica napusembryos. J. Biol. Chem. 278, 29442

    29453.

    Sims, A.P. and Folkes, B.F. (1964)A kinetic study of the assimilation

    of (15n)-ammonia and the synthesis of amino acids in an expo-

    nentially growing culture ofCandida utilis.Proc. R. Soc. Lond. B

    Biol. Sci.159, 479502.

    Soga, T., Ueno, Y., Naraoka, H., Ohashi, Y., Tomita, M. and

    Nishioka, T. (2002) Simultaneous determination of anionic inter-

    mediates for Bacillus subtilis metabolic pathways by capillaryelectrophoresis electrospray ionization mass spectrometry.Anal.

    Chem.74, 22332239.

    Stephanopoulos, G. N., Aristidou, A. A. and Nielsen, J. (1998)

    Metabolic Engineering: Principles and Methodologies. San Die-

    go, CA, USA: Academic Press.

    Voit, E. (2000) Computational Analysis of Biochemical Systems. A

    Practical Guide for Biochemists and Molecular Biologists.

    Cambridge, UK: Cambridge University Press.

    186 E. Heinzleet al.

    2007 The AuthorsJournal compilation 2007 Blackwell Publishing Ltd, The Plant Journal, (2007), 50, 176187

  • 8/14/2019 Estimation of Metabolic Fluxes

    12/12

    Wahl, S. A., Dauner, M. and Wiechert, W.(2004) New tools for mass

    isotopomer data evaluation in 13C flux analysis: mass isotope

    correction, data consistency checking, and precursor relation-

    ships. Biotechnol. Bioeng. 85, 259268.

    Wiechert, W.(2001) 13C metabolic flux analysis.Metab. Eng.3, 195

    206.

    Wiechert, W., Mollney, M., Petersen, S. and de Graaf, A. A. (2001) A

    universal framework for 13C metabolic flux analysis.Metab. Eng.

    3, 265283.Wittmann, C.(2002) Metabolic flux analysis using mass spectrom-

    etry.Adv. Biochem. Eng. Biotechnol.74, 3964.

    Wittmann, C. and Heinzle, E. (1999) Mass spectrometry for meta-

    bolic flux analysis.Biotechnol. Bioeng. 62, 739750.

    Wittmann, C. and Heinzle, E. (2001a) Application of MALDI-TOF MS

    to lysine-producing Corynebacterium glutamicum: a novel

    approach for metabolic flux analysis. Eur. J. Biochem.268, 2441

    2455.

    Wittmann, C. and Heinzle, E. (2001b) Modeling and experimental

    design for metabolic flux analysis of lysine-producing Coryne-

    bacteria by mass spectrometry.Metab. Eng.3, 173191.

    Method for metabolic flux and control analysis 187

    2007 The AuthorsJournal compilation 2007 Blackwell Publishing Ltd, The Plant Journal, (2007), 50, 176187