Vous êtes sur la page 1sur 15

68

Current Chemical Biology, 2008, 2, 68-82

Emerging Roles for Metabolic Engineering - Understanding Primitive and


Complex Metabolic Models and Their Relevance to Healthy and Diseased
Kidney Podocytes
Mehmet M. Altintas1, Kutlu O. Ulgen2, Darryl Palmer-Toy3, Vivian E. Shih4, Dhinakar S. Kompala5
and Jochen Reiser*,1
1

Division of Nephrology and Program in Glomerular Disease, Department of Medicine, Massachusetts General
Hospital and Harvard Medical School, Boston, MA 02129-2020, USA
2

Department of Chemical Engineering, Bogazici University, Bebek 34342 Istanbul, Turkey

Regional Reference Laboratories, Southern California Permanente Medical Group, North Hollywood, California
91605, USA
4

Amino Acid Disorders Laboratory, Massachusetts General Hospital and Harvard Medical School, Boston, MA 021292020, USA
5

Department of Chemical and Biological Engineering, University of Colorado, Boulder, Colorado 80309-0424, USA
Abstract: The central metabolism of a cell can determine its short- and long-term structure and function. When a disease
state arises, the metabolism (i.e., the transportation of nutrients into the cells, the overall substrate utilization and production, synthesis and accumulation of intracellular metabolites, etc.) is altered in a way that may permit organisms to survive
under the changing physiologic constraints. Although the response of cells to injury was studied thoroughly using molecular biology and structural morphology techniques, the knowledge regarding the metabolic signatures of the disease is limited. However, recent advances in analytical methods and mathematical tools have led to new approaches to those questions with the concept of computational biology which relies on the integration of experimentation, data processing and
modeling. The attempt to formulate current knowledge in mathematical terms has led to the development of several
mathematical modeling tools (i.e., metabolic flux analysis, metabolic control analysis, etc.) that helps us to understand an
entire biological system from basic structure to dynamic interactions. This review provides an overview and summarizes
the current status of applications of mathematical models for the quantification of fluxes. A specific example of kidney
podocyte cells illustrates how metabolic alterations, which occur during injury, can be used to aid in future therapeutic
development.

Keywords: Metabolic flux analysis, intracellular fluxes, podocyte metabolism, analysis of the health and disease.
INTRODUCTION
Biology in the past decades has been characterized by a
qualitative and descriptive approach designed to investigate
molecular behavior. Today, with computational technology
(i.e., modeling and simulation of complex processes), the
relationships between various parts of a biological system
(e.g., gene and protein networks involved in cell signaling,
metabolic pathways, organelles, cells, physiological systems,
organisms, etc.) are potentially understandable and predictive to some extent.
Together with the advances in biological science and
technology, an enormous number of new data types and formats are emerging daily. The success of mathematical models in biological and medicinal research requires an iterative
interaction between experimentation, modeling and simulation, and theory. These mathematical representations of
complex systems all have limitations, but they have the
*Address correspondence to this author at the Nephrology Division, Massachusetts General Hospital, Harvard Medical School, 149, 13th Street, Room
8214, Boston, MA 02129-2020, USA; Tel: 617-726-9363; Fax: 617-7265669; E-mail: jreiser@partners.org
1872-3136/08 $55.00+.00

potential to create a robust computational biology by the


collaborative work of biologists with mathematicians and
researchers from other disciplines [1].
Important progress has been made towards understanding
the complex systems in biological science from almost two
decades of research in metabolic engineering. The overall
goals of metabolic engineering are (i) to model and predict
the cellular metabolism in a quantitative manner and (ii) to
improve cellular properties by designing and implementing
rational genetic modifications [2-4]. In this sense, metabolic
engineering deals with the measurement of metabolic fluxes
and their control together with the evaluation of the impact
of genetic modifications on cellular physiology [5-8].
Based on this definition, metabolic engineering utilizes
various mathematical modeling tools (i.e., metabolic flux
analysis, metabolic control analysis, etc.) to identify fluxes
through critical metabolic pathways in the cell or tissue of
interest. Then, these quantitative methods allow precise control of the metabolic network throughout the pathway of interest and optimization of the metabolic flow to target metabolites or products. Besides providing a convenient framework for the integration of fluxes with the intracellular vari 2008 Bentham Science Publishers Ltd.

Emerging Roles for Metabolic Engineering

ables (metabolites, enzyme activities, proteins, etc.), metabolic engineering has a novel contribution with its capacity
to analyze more global networks of several enzymes and
reactions.
In this review, we describe the use of metabolic engineering techniques to analyze the physiology of normal and disease states in different model systems. We focus on the
pathway engineering approach towards novel therapies for
patients with injuries and/or chronic diseases, with an emphasis on disease processes occurring in podocytes, cells
essential to maintain kidney filtration.
MODELING METABOLIC NETWORKS
Mathematics and computation play critical roles in understanding the physiological behavior of cells and systematic integration of this information into a predictive model
that can be used for controlling the fate of the organism. In
this context, the essence of metabolic engineering is the capacity to engineer pathways that regulates the overall metabolism on the basis of a set of stoichiometric and/or kinetic
rules. In the following sections, mathematical tools of the
metabolic engineering will be discussed briefly before reviewing its applications to bacterial, yeast and mammalian
systems.
Construction of the Metabolic Networks
A metabolic pathway is defined as the series of feasible
and observable biochemical reactions connecting a specified
set of input and output metabolites [7]. A metabolic pathway
may be linear, cyclic, branched, tiered, directly reversible, or
indirectly reversible. Various metabolic pathways within a
cell (glycolysis, tricarboxylic acid cycle, pentose phosphate
pathway, gluconeogenesis, glyoxylate shunt, oxidative phosphorylation, etc.) form the cells metabolic network that is
generally a complex and nonlinear system of cellular (metabolites, nucleic acids, etc.) and extracellular constituents
(substrates, protons, etc.) and reactions. The metabolic networks including central carbon metabolism, regulation
mechanisms, energetic and transport reactions have a key
role in sustaining cellular functions by coordinating the activity of different metabolic pathways [9-11].
Since the metabolic pathways and fluxes are at the core
of metabolic modeling, mapping biochemical networks in
the cell or organ of interest is the priority for the application
of metabolic engineering. Those networks are currently organized into databases such as KEGG [12] and MetaCyc
[13]. These databases are based on information from experimental data and store valuable information about hundreds of pathways and cellular processes. In addition, the
complete metabolic network may not be fully described, i.e.,
pathway(s) within the network may consist of alternative
reaction(s) that produce the same set of metabolites from the
same set of precursor metabolites and cofactors [9]. Therefore, the network map is refined in an iterative fashion for
the most accurate reflection of the existing biochemical
knowledge (Fig. 1).
Measurement of Fluxes
Metabolic flux is the rate of material that flows along a
metabolic pathway or even through a single reaction connecting two or more metabolites. Measurement of metabolic

Current Chemical Biology, 2008, Vol. 2, No. 1

69

fluxes is an important quantitative approach in metabolic


engineering due to its capacity to characterize the flux distributions, reaction mechanisms and associated parameters.
It is possible to determine the intracellular fluxes by using isotope tracer methods. In isotopomer analysis, the biological system (of the microorganism) is fed with a specifically labeled substrate (usually 13C or 15N) during the stationary growth phase and the labeling patterns of the compounds in the central C or N metabolism (e.g., isotopic enrichment in the intracellular metabolite pools) are measured
after the labeling isotopic balances are reached. This technique is either based on nuclear magnetic resonance or mass
spectrometry [14-20]. The resulting data, together with the
metabolite balancing, provide a large amount of additional
information regarding pathway identification (i.e., identification of active pathways, analysis of cyclic, parallel and split
pathways, etc.), flux distribution and subcellular compartmentation.
Analysis of Fluxes by Mathematical Tools
Intracellular fluxes can also be determined from the
measured extracellular metabolite fluxes (e.g., the fluxes of
substrates into the cells and the fluxes of metabolites that are
secreted out of the cells) using mathematical tools such as
biochemical systems theory [21-25], metabolic control
analysis [26-29], metabolic flux analysis [30-35] and cybernetic modeling [36-42]. With the exception of metabolic flux
analysis (MFA), these approaches require information on the
kinetics of the individual reactions although the level of detail in the functional form of kinetics may be rather low.
The computational study of metabolic systems is a subject with a long history [43]. Mathematical methods for sensitivity analysis of metabolic regulation began in the late
1960s [44-46] and led to the biochemical systems theory
(BST). All processes are represented as products of powerlaw functions (either S-systems or the alternative variant of
generalized mass action systems, GMA) in BST to investigate metabolic and gene-regulatory systems. A major advantage of the BST approach is that it does not require a priori
structural information on the underlying pathway and models
are designed based solely on the identity of the reactants and
their reactional and regulatory interconnections.
The metabolic control analysis (MCA) shares the same
fundamental definition with BST. Both methods use the
same experimental information to study the responses of
metabolic systems to changes in their parameters that can be
exerted by genetic manipulations, enzymatic titrations or
enzymatic inhibitors. MCA (also known as metabolic control
theory) assumes that there is a definite amount of flux control that emerges as a parameter of critical importance in the
identification of feasible enzymatic modifications having
maximal impact on the network flux. In metabolic systems,
this control is normally spread quantitatively among several
enzymes in the system. Since its introduction three decades
ago, MCA has received much theoretical and experimental
attention since it allows us to understand how metabolic
fluxes are controlled by certain enzyme activities and metabolite concentrations [47].
Assuming that cellular metabolism is at a steady state,
MFA (also referred to as flux balance analysis, FBA) allows
us to calculate the unknown internal fluxes over each reac-

70 Current Chemical Biology, 2008, Vol. 2, No. 1

Altintas et al.

Fig. (1). The necessary steps to construct a metabolic model.

tion using data on external fluxes, metabolic stoichiometry


and mass balances around intracellular metabolites. Besides
quantification of pathway fluxes, MFA is useful for the identification of the rigidity (or flexibility) of key metabolic
branch points, the determination of non-measured extracellular fluxes, the calculation of maximum theoretical yields, the
identification of alternative pathways and the observation of
the function of metabolic pathways in vivo [6, 48, 49]. Although MFA has the ability to estimate the unmeasured internal fluxes, it does not incorporate any kind of regulation
mechanisms. On the other hand, MFA can incorporate additional information when it is available and is a powerful
technique when used in conjunction with novel physiological
experiments.
Another powerful methodology for describing the complex phenomena observed in biological cells is the cybernetic modeling framework which was developed by Ramkrishna and coworkers [50-52]. The cybernetic approach is
based on the idea of proposing an optimal mechanism such
that the cell regulates the synthesis and activity of the key
enzymes to promote perceived local or global cellular objective. The application of cybernetic framework to metabolic
engineering requires limited kinetic information combined
with cybernetic variables that modify the rates of enzyme
production and activation to modify the metabolic reaction
rates. The definition of the cybernetic variables varies depending on the nature of the metabolic pathway being examined [40].
In the presence of detailed kinetic information about the
specific cellular processes (e.g. enzyme catalyzed reactions,
protein-protein interactions, or protein-DNA binding), it is

possible to combine kinetics with the known stoichiometry


of the model to interpret and predict cellular behavior [25,
53, 54]. In particular, a kinetic model is constructed by (i)
defining all reactions with their stoichiometry, (ii) specifying
the kinetics of each reaction, and (iii) setting up the proper
mass balances for each metabolite. One major drawback of
kinetic modeling is that a realistic kinetic modeling of metabolic networks needs mechanistically-correct rate equations
that require detailed knowledge of all physiological effectors
influencing the activity of the catalyzing enzyme. Moreover,
our biochemical knowledge is limited to the enzymatic properties in vitro which are very different from the conditions
inside the microorganism, in vivo. Therefore, kinetic modeling is not suitable for developing metabolic engineering
strategies in the absence of well-characterized kinetic parameters [55, 56]. On the other hand, once kinetically judged
and experimentally validated, these models replace timeconsuming and expensive experiments for accurate prediction of the fluxes and form a valuable basis for viewing the
whole pathway, which adds to our understanding of the system of interest [57].
Beyond their intrinsic differences, all these modeling
approaches provide a mathematical basis for the representation of the existing relationship between metabolic network
components (metabolites, enzymes, etc.) and the cellular
systemic behavior exhibited by the network of interest. As
metabolite levels are governed by changes in fluxes and enzyme activities, understanding the changes on a metabolic
level will then help in understanding the regulatory mechanisms that lead to these changes in fluxes [58]. The metabolic model can thus form a solid biochemical basis on com-

Emerging Roles for Metabolic Engineering

puting the implications of the functions of the cell, such as


signaling and regulatory networks.
Despite the enormous potential and success of these approaches, the components of a system and their interactions
in the cellular network need to be studied further in order to
have a system-wide picture and system-level understanding
of biology [59, 60]. It is important as many properties of life
arise at the systems level only. More recently, systems biology has emerged as an integrative approach that investigates
pathways and networks by combining data about genes, proteins, enzymes and metabolites to generate a comprehensive
picture of the system (e.g., tissue, organ or organism) as a
whole [61-71].
Because of the biological complexity, system biology
enables a wide spectrum of mathematical modeling techniques such as the ones discussed above and bioinformatics
to visualize large networks of cellular components and generate hypotheses about the cell behavior. These hypotheses
can then be tested experimentally in systems biology framework by integrating the detailed information about the entire
genome of an organism (genomics), small molecules that
cells assimilate or synthesize (metabolomics), cellular proteins and their physical interactions (proteomics), the temporal and spatial distribution of all gene transcripts (transcriptomics) and the data from other accompanying highthroughput experiments.
METABOLIC MODELING OF HEALTHY AND DISEASED PODOCYTES
Podocytes, also known as the visceral glomerular epithelial cells, are terminally differentiated cells with a complex
cellular architecture contributing to many functions of the

Current Chemical Biology, 2008, Vol. 2, No. 1

71

normal kidney glomerulus [72]. They possess a highly


branched array of actin-rich foot processes (FP) that are essential to glomerular filtration on the kidney. The FPs of
neighboring podocytes regularly interdigitate, leaving between filtration slits that are bridged by an extracellular
structure, known as the slit diaphragm (SD). The FPs are
anchored on the outer aspect of the glomerular basement
membrane (GBM) (Fig. 2). Therefore, podocytes define the
final barrier to protein loss, which explains why podocyte
injury is typically associated with marked urinary protein
loss (proteinuria) which is a risk factor for morbidity and
mortality in these patients. Characteristically, podocyte FPs
are rearranged upon injury leading to loss of FP interdigitations and forming a flattened sheet of cytoplasm above the
GBM (referred to as FP effacement). If FP effacement is
reversed, coordinated kidney filtration is re-established.
However, in many cases, FP effacement progresses into
more severe and irreversible podocyte injury resulting in
progressive kidney failure. The metabolic signatures that are
involved to maintain normal and diseases podocyte structures need to be defined to better understand and treat proteinuric kidney disease.
Podocytes are terminally differentiated cells and their
metabolism is complicated and its understanding is just beginning. These complexities render the material balances for
podocyte cell cultivation a difficult problem to solve. Nevertheless, detailed material balance analysis in podocyte cell
culture systems would be desirable to provide insights toward a better understanding of podocyte cell metabolism.
One means for characterizing the intracellular metabolism of
cultured podocytes under normal and disease conditions is
the identification of the flux distributions by MFA as it

Fig. (2). Podocytes are highly specialized cells within the glomerulus that are essential for ultrafiltration. They form foot processes (FP) that
are highly dynamic cellular extentions. FPs, that rest on glomerular basement membrane (GBM), are interconnected by the slit diaphragms
(SD) to form the final component of the kidney permeability barrier.

72 Current Chemical Biology, 2008, Vol. 2, No. 1

Table 1.

Altintas et al.

Application of MFA to Different Microorganisms


Organism

Bacterium
Bacillus licheniformis
Bacillus subtilis

Clostridium acetobutylicum
Clostridium cellulolyticum
Corynebacterium glutamicum

Corynebacterium melassecola
Escherichia coli

Lactobacillus lactis
Streptomyces coelicolor
Streptomyces lividans
Zymomonas mobilis
Yeast
Aspergillus niger
Aspergillus oryzae
Candida milleri
Candida tropicalis
Penicillium chrysogenum
Saccharomyces cerevisiae

Saccharomyces kluyveri
Mammalian cells
BHK cells
CHO cells
HEK 293 cells

Hybridomas

Organ systems
Adipose tissue
Brain
Heart
Liver

Skeletal muscle

MFA Application

Ref.

Serine alkaline protease fermentation


Analysis of the energetic efficiency of growth
Glucose metabolism
Riboflavin production
Acetone fermentation
Fermentation characterization of the recombinant strains
Cellulose metabolism
Glutamate production
Lysine production from glucose
Lysine production from glucose and fructose
Lysine production from sucrose
Metabolism of mutant cells
Glucose and fructose metabolism
Acetate secretion under ATP maximization conditions
Byproduct secretion under various oxygenation rates
Effects of gene additions or deletions on metabolism
Effects of plasmid maintenance on metabolism
Growth and metabolism
Polyphosphate metabolism
Succinic acid/succinate production
Lactic acid production
Actinorhodin production
Calcium dependent antibiotic (CDA) production
Growth and metabolism
Glucose, fructose and xylose metabolism

[134]
[135]
[136]
[137-140]
[141-143]
[144]
[145, 146]
[147-149]
[150-157]
[158]
[159]
[160, 161]
[162]
[163, 164]
[165]
[166-175]
[176]
[177-181]
[182]
[183-185]
[186]
[187]
[188]
[189, 190]
[191]

Glucose and glutamate metabolism


Nitrogen metabolism
Xylose metabolism
Dicarboxylic acid (DCA) production
Xylose metabolism
Penicillin production
Glucose metabolism and ethanol production
Growth on different carbon sources
Xylose metabolism and ethanol production
Nitrogen metabolism and ethanol production
Starch metabolism and ethanol production
Glycerol production
Heterologous protein production
Glucose metabolism

[192]
[193]
[194]
[195]
[196]
[197-200]
[201-206]
[207]
[205, 208-213]
[214]
[215]
[216]
[217]
[218]

Glucose and glutamine metabolism


Glycosylation
Growth and metabolism
Adenoviral vector production
Effects of genetic engineering on metabolism
Glucose and glutamine metabolism
Energy metabolism
Glutamate production
Growth and metabolism
Medium design and optimization
Response to changing medium composition

[219]
[220]
[111, 221, 222]
[223]
[113, 132]
[112]
[224]
[225]
[108-110, 226-228]
[229-233]
[107, 234]

Adipocyte formation
Fat synthesis
Glutamate metabolism
Neuron-astrocyte coupling
Characterization of the human mitochondrial metabolic network
Mitochondrial network properties in health and disease
Analysis of the hypermetabolic state
Collagen synthesis
Hepatocyte function in plasma
Interhepatic metabolism in liver failure
Postburn hepatic metabolism
Analysis of the postburn hypermetabolic state

[235]
[236]
[237]
[238]
[239]
[240]
[241-243]
[244]
[114, 245]
[115, 246]
[247-249]
[250]

Emerging Roles for Metabolic Engineering

offers the advantage of simplicity, i.e., it solely relies on the


known stoichiometry of a given biochemical reaction network. This approach has been employed successfully to analyze the metabolic physiology and behaviors of a variety of
species, from single-celled to multicellular organisms with
complex organ systems (Table 1).
Response of Podocytes to Injury
Podocytes are injured in many forms of human and experimental glomerular disease [73-76], including minimal
change disease (MCD), focal segmental glomerulosclerosis
(FSGS), diabetes mellitus (DM), membranous glomerulopathy (MG), crescentic (rapidly progressive) glomerulonephritis (CGN, RPGN) and lupus nephritis (LN). The early events
are characterized by alteration in SD and FP configuration,
resulting in FP effacement and loss of SD integrity. FP effacement is associated with the onset of proteinuria and is
accompanied by a reorganization of the actin cytoskeleton
into a dense network [77]. These early changes are fully reversible [78, 79].
In 1957, Farquhar et al. [80] were the first to describe
extensive FP effacement in biopsies of patients with nephrotic syndrome. Since then, FP effacement has been subject
of extensive investigation in humans and in animal models.
Based on recent insight into the molecular pathology of
podocyte injury, at least six major causes of FP effacement
and proteinuria can be identified: (i) interference with the
negative surface charge of podocytes [81-84] or slit membrane [85, 86]; (ii) interference with the activity of GLEPP1,
a membrane bound tyrosine phosphatase [87]; (iii) interference with the actin cytoskeleton and its associated protein
alpha-actinin-4 [88, 89]; (iv) interference with the GBM or
the podocyte-GBM interaction [90-97]; (v) interference with
the SD complex [98-100] and (vi) induction of danger signals in podocytes via upregulation of podocyte B7-1 [101].
Disease Models Under Investigation
In order to analyze the relationship of podocyte structure
and their central metabolism and the pathogenetic mechanisms of podocyte FP effacement, we utilize three in vitro
disease models: (i) podocytes with alpha3 integrin deletion
(genetic model [102]), (ii) podocytes treated with lipopolysaccharides (LPS, immunological model [103]), (iii) podocytes treated with puromycin aminonucleoside (PAN, toxic
model [104]). The common denominator of all these cell
models is the rearrangement of the actin cytoskeleton.
The deletion of alpha3 integrin in mice leads to severe
podocyte FP effacement and perinatal death of mutant mice
from kidney and lung failure [91]. We have developed conditionally immortalized mouse podocyte cell lines from these
mice. These cells have functional and structural deficits and
serve as a genetic in vitro model of podocyte damage [102].
In a recent study, we have established that these cells have
an induced expression of the protease cathepsin L. Moreover, these cells exhibit different cytoskeletal dynamics, e.g.
increased cellular motility when compared to wildtype podocytes.
Similar structural changes as seen in alpha3 integrin deficient podocytes can be achieved by treating wildtype podocytes with PAN. PAN treatment leads to the reorganization
of the actin cytoskeleton, focal contacts, and SD proteins in

Current Chemical Biology, 2008, Vol. 2, No. 1

73

cultured podocytes [104]. We also observed that application


of PAN to wildtype podocytes for 48 hours caused a dosedependent increase in autologous B7-1 expression, a transmembrane protein previously known as a B-cell costimulatory molecule. B7-1 in podocytes is capable to induce FP
effacement and disruption of the SD complex, thereby modifying glomerular permselectivity [103]. In another work, we
reported that PAN treatment strongly increased the migration
of wildtype podocytes [102].
The application of LPS to cultured podocytes represents
a more recent model. Podocytes sense LPS by expressing
toll-like receptor-4 (TLR-4) and its coreceptor CD14, which
in turn triggers B7-1 expression and the reorganization of the
podocyte actin cytoskeleton. Most significantly, LPS injection caused podocyte FP effacement and severe proteinuria
within 24 hours when injected into mice. The LPS-induced
proteinuria is transient and returns to base-line levels after 72
hours [103]. In summary, the above models have similar
structural deficits in common but the underlying pathways
are diverse. These models are therefore well suited to dissect
different metabolic profiles for a similar appearing structural
disease condition.
Podocyte Metabolic Network
Traditional MFA starts with a predetermined metabolic
network. Although the central metabolic pathways are well
known, the network models used in the literature differ due
to the simplifications made in the case of non-identifiable
fluxes. Those simplifications are made by either lumping the
reactions in a pathway that can not be determined independently or assuming unidirectionality (e.g., zero exchange flux)
in some fluxes, generally the ones belonging to product formations [105]. Also, there are cases where a pathway or reaction may assumed to be inactive or not expressed [106].
The metabolic network model for podocytes (Fig. 3), is a
moderately detailed model that we use as a basis for quantification of the fluxes for the cultured podocyte cells. It is an
adaptation of a network previously used for hybridomas
[107-110], CHO cells [111], HEK 293 cells [112, 113] and
hepatocytes [114]. In contrast to CHO cells, HEK 293 cells
and hepatocytes, podocytes represent highly differentiated
cells. This means that the metabolism of podocytes is tailored to maintain such a differentiated phenotype and may
quantitatively differ from proliferative cells such as HEK
293 cells. On the other hand, the operative metabolic enzyme
composition is likely similar between various mammalian
cell types, so that we decided to use available information of
mammalian cells to construct our podocyte metabolic network. This network-adaptation approach was employed successfully to analyze the podocyte behavior by manually adding or removing the corresponding pathways in the metabolic map.
In the model, there are 40 intracellular fluxes and 24
fluxes for transport rates and biosynthesis rates that can be
measured. The transport fluxes are formally defined for each
measured metabolite (all 20 amino acids, glucose, lactate,
ornithine and citrulline), and each rate is defined with a positive sign for production. Each extracellular metabolite is
linked to its intracellular counterpart metabolite pool. Sixtythree metabolites constitute the nodes for pseudo steady-state
mass balances. Intracellular fluxes are overall biochemical

74 Current Chemical Biology, 2008, Vol. 2, No. 1

Altintas et al.

Fig. (3). The metabolic network model for podocyte cultures. Arrows indicate the direction of reaction.

reactions representing kidney-specific functions and pathways: glycolysis (flux nos. 1 to 5), reduction of pyruvate to
lactate (no. 6), Krebs cycle (nos. 7 to 14), urea cycle (nos. 15
to 17), amino acid degradation (nos. 18 to 37), pentose phosphate pathway (no. 38), oxygen uptake and electron transport
(nos. 39 and 40) (Fig. 3). All pathways were verified as feasible for Mus musculus by online bioinformatic databases
[12, 13].
Formulation of the Model
MFA starts with setting up a stoichiometric matrix for a
network of reactions occuring in the cell. Then, the mass
balance constraints around intracellular metabolites are
specified. These constraints identify a series of linear equations of individual reaction fluxes that must be fulfilled to
enable steady state criterion. Mathematically, the reaction

network and the mass balance constraints can be summarized


in the following matrix notation:

 S11 S12
S  r =  :
:
Sm1 Sm2

   S1n 
 r1 

:
:    :  = 0
   Smn  mxn  rn  nx1

(Eq. 1)

S is the mxn stoichiometry matrix, with m as the number


of metabolites, and n as the number of reactions. The vector
r represents all the individual fluxes of intracellular and extracellular compounds (Fig. 4). In this equation, S11 denotes
the stoichiometric coefficient of the first metabolite in the
first reaction while S12 denotes for the stoichiometric coefficient of the first metabolite in the second reaction, etc. For
example, if the third metabolite participates only in its formation in reaction 3 and its disappearance in reaction 7 with

Emerging Roles for Metabolic Engineering

Current Chemical Biology, 2008, Vol. 2, No. 1

75

Fig. (4). Principles of metabolic flux analysis (MFA). MFA technique is based on relatively simple linear algebra. If the stoichiometry of the
relevant intracellular reactions and the cellular composition are known, and the uptake and secretion rates of the relevant metabolites (rates
denoted by e in the figure) have been measured, the reaction rates (rates denoted by r in the figure) can be determined using the appropriate mass balance equations.

the stoichiometric coefficient of 1, then S33 = -S37 = 1, while


all other coefficients in that row are zero. The right hand side
of this equation is made equal to zero assuming the cultured
podocytes are in metabolic steady state where the intracellular levels of metabolites are constant [48].
By separating r into measured and unknown components,
and rcalc, respectively, and partitioning matrix S into
r
meas
and Scalc, where they contain the stoichiometric coeffiS
cients of measured and unknown reactions (i.e., internal and
transport fluxes), respectively, we obtained:
meas

Scalc x rcalc + Smeas x rmeas = 0

(Eq. 2a)

and
Scalc x rcalc = Smeas x rmeas

(Eq. 2b)

calc

is not a square matrix, i.e., the number of


Because S
rows is greater than the number of columns, Eq. (2b) cannot
be solved by simple inversion. One approach was to use the
Moore-Penrose pseudo-inverse method [48], in which each
side of Eq. (2b) is multiplied by the transpose of Scalc:
(Scalc)T x Scalc x rcalc = (Scalc)T x Smeas x rmeas
calc

The matrix multiplier of r (i.e., (S


invertible, and Eq. (3) is easily solved:

calc T

) xS

(Eq. 3)
calc

) is now

rcalc = ((Scalc)T x Scalc)-1 x (Scalc)T x Smeas x rmeas

(Eq. 4)

Since the system of linear equations is overdetermined


(more equations than unknown fluxes), fluxes are determined by linear regression [48, 115].
Proteome Analysis to Complement MFA
Proteomics is one among various omics fields that allows the analysis of complex mixtures of proteins, for example from cell lysates, subcellular fractions or immunoprecipitated protein complexes [116-118]. The general approach to
cellular proteomics is characterized by (i) protein extraction
from a crude homogenate, (ii) separation of the intact proteins (e.g. by 2D-PAGE) or peptide fragments of specific
proteolytic digestion (e.g. by trypsin digestion followed by
liquid chromatography) and (iii) identification of these proteins by mass spectrometry and bioinformatics [119]. Quantitative comparisons of two or more proteomes are possible
if the protein samples are distinctly labeled, stoichiometrically combined, and the mixture is then separated and analyzed [120, 121]. While 2D-PAGE has been the central technology in proteomics for over 30 years, there have been tremendous developments in liquid chromatography-mass spectrometry (LC-MS) to combine separation and analysis into a

76 Current Chemical Biology, 2008, Vol. 2, No. 1

single on-line process capable of the rapid identification of


hundreds or thousands of proteins from cells [122].
Ransom [123] attempted to construct a whole cellular
proteome map for podocyte cells for the first time. He and
his coworkers performed a differential analysis using 2DPAGE of changes in protein expression in cultured murine
podocytes in response to glucocortioids [124]. A total of 106
proteins representing 88 unique proteins were identified in
their podocyte proteome map.
To elucidate the molecular basis of biological processes,
the analysis of dynamic changes of the subcellular distribution of proteins is necessary [125]. In order to describe the
subcellular proteome of the podocyte and to identify podocyte proteins whose expression is altered by toxic challenges, cytoplasmic andmitochondrial proteome analysis of
wildtype and disease models of cultured podocytes were
performed by LC-MS. The results are summarized in Fig. 5.
Organellar proteomics is a valuable tool to complement and
substantiate MFA studies in cultured cells, i.e. podocytes.
Comparison of organellar proteomes in healthy and diseased
podocyte cells will complement metabolic studies and provide further insights into disease pathogenesis. Such a comparison will also lay ground for additional functional studies
to prove or adapt results obtained by MFA. In particular,
quantitative changes in proteome permit us understand
measureable changes in metabolite fluxes at the enzyme
level.
Evaluation of the Model
In batch cultures of diseased podocytes, we observed an
increased use of both glucose and glutamine when compared
to the healthy cells. An increase in the oxidation of glucose
and glutamine in diseased cells promoted higher utilization
of other amino acids as well. MFA showed that essential
amino acids contributed to about 80-90% of the total amino
acid consumption and glutamine is needed to supply the essential amino acids in diseased cells. On the other hand, the
nitrogen-balance calculations revealed that the amount of
nitrogen incorporated by podocytes (mainly from glutamine
uptake) exceeded significantly the amount of nitrogen removed as glutamate, alanine and aspartate in disease models.
The ammonia measurements in disease podocytes correlated
very well with the nitrogen-balance calculations suggesting
that a considerable part of nitrogen is released in the form of
ammonia. This ammonium buildup as a result of high glutamine consumption in diseased podocytes makes sense
since glutamine metabolism and glutamine chemical decomposition are the primary sources of ammonium buildup in the
cell cytoplasm and the culture media [126] and specific production rate of ammonium is closely correlated to glutamine
[127]. Schneider et al. [128] summarized a large number of
articles dealing with numerous different effects caused by
elevated ammonia and ammonium concentrations on cell
cultures. The challenge here is to generate or employ a metabolic engineering tool to help diseased cells alleviate ammonia stress.
High glucose and glutamine uptake and the large amount
of lactate release by diseased podocyte cultures prompted us
to think about the possibility that these cells consume
amounts of glucose and glutamine higher than that required
to satisfy their energy and biosynthetic needs which may

Altintas et al.

result in an accumulation of lactate and ammonia in the medium. If the metabolism of glucose or glutamine can be improved, then it is possible to obtain higher consumption rates
of these nutrients and lower lactate and ammonia production.
It is also known that there is a close interrelation of the consumption of these two nutrients by the cells to satisfy their
carbon and nitrogen demands [129-132]. One way to direct
carbon sources effectively through pathways other than the
lactate production pathway is to promote the conversion of
pyruvate to acetyl-CoA and CO2 via pyruvate dehydrogenase
(Fig. 3). This branching enzyme from glycolysis into the
TCA cycle is not active in mammalian cell lines [133] as
validated by our proteome analysis. The stable expression of
pyruvate dehydrogenase in podocyte cells would improve
the utilization of glucose and limit the production of lactate
and ammonia, which are deleterious to the cell.
In mammalian cells, the removal of amino nitrogen from
the glutamate family of amino acids (arginine, ornithine,
proline, histidine and glutamine) is achieved through a welldescribed transdeamination system involving aspartate
transaminase (catalyzing the transfer of an amino group from
glutamate to oxaloacetate, forming alpha-ketoglutarate and
aspartate) and glutamate dehydrogenase (catalyzing the reversible oxidative deamination of glutamate to alphaketoglutarate and ammonia). Proteomic analysis revealed
that these two enzymes were down-regulated in PAN-treated
cells indicating the possibility of ammonia accumulation in
these cells which is also pointed out by the comparative
MFA on untreated (control model) and PAN-treated (toxic
model) cells (Fig. 5).
CONCLUSION
The glomerular podocyte is increasingly recognized as a
primary determinant of many glomerular diseases (e.g., focal
segmental glomerular sclerosis, glomerulonephritis, membranous nephropathy, glomerular hypertension, etc.) leading
to chronic kidney disease. Recent discoveries have highlighted the importance of podocyte proteins in maintaining
the glomerular filtration barrier in both health and disease.
Although the response of podocytes to injury was studied
thoroughly using molecular biology and structural morphology research tools, the metabolic signatures of podocyte FP
effacement are not known. Metabolic analysis with a clear
correspondence between disease and the related metabolic
changes by employing metabolic flux analysis (MFA) and
proteomic analyses help to define important metabolic pathways in healthy and diseased podocytes and can also be employed for other eukaryotic disease models.
Although MFA has been applied extensively to study the
cellular function (i.e., to increase the production of the primary or secondary metabolites, to produce desired chemicals
from less expensive feedstocks, to generate alternative pathways for the production of specialty chemicals, etc.) from
bacteria, yeasts, fungi, and mammalian cells, its application
to problems in physiology and medicine is less well recognized [6]. Here, we developed a thorough stoichiometric
model and used MFA to obtain a quantitative estimate of
stationary metabolic flux rates without knowledge of the
detailed kinetics of individual reactions, which are usually
accompanied with problems concerning availability and in
vivo/in vitro discrepancies. The proposed metabolic model-

Emerging Roles for Metabolic Engineering

Current Chemical Biology, 2008, Vol. 2, No. 1

77

Fig. (5). Relative abundances of some key enzymes in the podocyte metabolic network in health and disease. Subcellular proteome data were
obtained from wildtype and disease (toxic) models of cultured podocytes, respectively.

ing has the capacity to determine the metabolic capabilities


of podocytes under normal and disease conditions by employing the framework of metabolite balancing in combination with linear programming methods. The changes observed in the profiles of glucose, glutamine and ammonia
give credence to the indication that not only the structure
(e.g., morphology) but also the metabolism of podocyte cells
is changed in disease.

Reiser). J. Reiser was also supported by the KMD foundation. M. M. Altintas was supported by an NIH training grant
and T32DK007540.
REFERENCES
[1]
[2]

We also performed a differential proteomic analysis of


cultured podocyte cells to better understand the physiology
under disease. The combinatory use of MFA and proteomic
analysis, the morphological differences between healthy and
diseased cells can be analyzed in detail and physiological
inheritance of these cells can be represented on metabolic
and proteome maps, respectively.

[3]

These findings will ultimatively lead to better understanding of kidney disease and help to refine pharmacological interventions. In addition, they may contribute to identification and characterization of potential new therapeutic
targets, biomarker discovery, prediction of therapeutic response, better therapeutic outcome and ultimately prevention
or treatment of the disease.

[7]

ACKNOWLEDGEMENTS
This work was supported by the American Society for
Nephrology (to J. Reiser) and the NIH (R01 DK073495 to J.

[4]

[5]
[6]

[8]
[9]

[10]
[11]

Csete ME, Doyle JC. Reverse engineering of biological complexity. Science 2002; 295: 1664-1669.
Bailey JE. Towards a science of metabolic engineering. Science
1991; 252: 1668-1674.
Stephanopoulos G. Metabolic engineering. Curr Opin Biotech
1994; 5: 196-200.
Nielsen J. Metabolic engineering: Techniques for analysis of targets for genetic manipulations. Biotechnol Bioeng 1998; 58: 125132.
Bailey JE. Lessons from metabolic engineering for functional genomics and drug discovery. Nat Biotechnol 1999; 17: 616-618.
Koffas M, Roberge C, Lee K, Stephanopoulos G. Metabolic engineering. Annu Rev Biomed Eng 1999; 1: 535-557.
Stephanopoulos G. Metabolic fluxes and metabolic engineering.
Metab Eng 1999; 1: 1-11.
Stephanopoulos G, Stafford DE. Metabolic engineering: A new
frontier of chemical reaction engineering. Chem Eng Sci 2002; 57:
2595-2602.
Jeong H, Tombor B, Albert R, Oltavi ZN, Barabasi AL. The largescale organization of metabolic networks. Nature 2000; 407: 651654.
Wagner A, Fell D. The small world inside large metabolic networks. Proc R Soc Lond 2001; 268: 1803-1810.
Hatzimanikatis V, Li C, Ionita JA, Broadbelt LJ. Metabolic networks: Enzyme function and metabolite structure. Curr Opin Struct
Biol 2004; 14: 300-306.

78 Current Chemical Biology, 2008, Vol. 2, No. 1


[12]

[13]
[14]

[15]
[16]
[17]
[18]
[19]

[20]
[21]
[22]

[23]
[24]

[25]
[26]
[27]

[28]
[29]
[30]
[31]
[32]

[33]
[34]

[35]
[36]
[37]
[38]

[39]

Kanehisa M, Goto S, Kawashima S, Okuno Y, Hattori M. The


KEGG resource for deciphering the genome. Nucleic Acids Res
2004; 32: D277-D280.
Krieger CJ, Zhang PF, Mueller LA, et al. MetaCyc: A multiorganism database of metabolic pathways and enzymes. Nucleic Acids
Res 2004; 32: D438-D442.
Wiechert W, de Graaf AA. In vivo stationary flux analysis by 13C
labeling experiments. Adv Biochem Eng Biotechnol 1996; 54: 109154.
Szyperski T. 13C-NMR, MS and metabolic flux balancing in biotechnology research. Quart Rev Biophys 1998; 31: 41-106.
Christensen B, Nielsen J. Isotopomer analysis using GC-MS. Metab Eng 1999; 1: 282-290.
Kelleher JK. Flux estimation using isotopic tracers: Common
ground for metabolic physiology and metabolic engineering. Metab
Eng 2001; 3: 100-110.
Wiechert W. 13 C metabolic flux analysis. Metab Eng 2001; 3: 195206.
Christensen B, Gombert AK, Nielsen J. Analysis of flux estimates
based on 13C-labelling experiments. Eur J Biochem 2002; 269:
2795-2800.
Wittmann C. Metabolic flux analysis using mass spectrometry. Adv
Biochem Eng Biotechnol 2002; 74: 39-64.
Savageau MA, Voit EO, Irvine DH. Biochemical systems theory
and metabolic control theory: 1. Fundamental similarities and differences. Math Biosci 1987; 86: 127-145.
Savageau MA, Voit EO, Irvine DH. Biochemical systems theory
and metabolic control theory: 2. The role of summation and connectivity relationships. Math Biosci 1987; 86: 147-169.
Savageau MA. Biochemical systems theory: Operational differences among variant representations and their significance. J Theor
Biol 1991; 151: 509-530.
Voit EO. Computational analysis of biochemical systems. A practical guide for biochemists and molecular biologists. Cambridge
University Press, Cambridge, UK, 2000.
Torres NV, Voit EO. Pathway analysis and optimization in metabolic engineering. Cambridge University Press, Cambridge, UK,
2002.
Kacser H, Burns JA. The control of flux. Symp Soc Exp Biol 1973;
27: 65-104.
Kell DB, Westerhoff HV. Metabolic control theory: Its role in
microbiology and biotechnology. FEMS Microbiol Rev 1986; 39:
305-320.
Kacser H, Burns JA, Fell DA. The control of flux. Biochem Soc
Trans 1995; 23: 341-366.
Fell D. Understanding the control of metabolism. Portland Press,
London, UK, 1996.
Varma A, Palsson BO. Metabolic flux balancing: Basic concepts,
scientific and practical use. Bio-Technol 1994; 12: 994-998.
Bonarius HPJ, Schmid G, Tramper J. Flux analysis of underdetermined metabolic networks: The quest for the missing constraints.
Trends Biotechnol 1997; 15: 308-314.
Lee K, Berthiaume F, Stephanopoulos GN, Yarmush ML. Metabolic flux analysis: A powerful tool for monitoring tissue function.
Tissue Eng 1999; 5: 347-368.
Christensen B, Nielsen J. Metabolic network analysis: A powerful
tool in metabolic engineering. Adv Biochem Eng Biotechnol 2000;
66: 209-231.
Cornish-Bowden A, Cardenas ML. From genome to cellular phenotype: A role for metabolic flux analysis? Nat Biotechnol 2000; 18:
267-268.
Wiechert W. Modeling and simulation: Tools for metabolic engineering. J Biotechnol 2002; 94: 37-63.
Baloo S, Ramkrishna D. Metabolic regulation in bacterial continuous cultures-I. Biotech Bioeng 1991; 20: 1337-1352.
Straight J, Ramkrishna D. Complex growth dynamics in batch
cultures: Experiments and cybernetic models. Biotech Bioeng
1991; 37: 895-909.
Straight J, Ramkrishna D. Cybernetic modeling and regulation of
metabolic pathways: Growth on complementary nutrients. Biotechnol Prog 1994; 10: 574-587.
Varner J, Ramkrishna D. Application of cybernetic models to
metabolic engineering: Investigation of storage pathways. Biotechnol Bioeng 1998; 58: 282-291.

Altintas et al.
[40]

[41]

[42]
[43]
[44]

[45]
[46]

[47]
[48]
[49]

[50]
[51]

[52]
[53]

[54]
[55]
[56]

[57]
[58]

[59]

[60]
[61]

[62]
[63]
[64]
[65]
[66]
[67]
[68]

Varner J, Ramkrishna D. Metabolic engineering from a cybernetic


perspective: I. Theoretical preliminaries. Biotechnol Prog 1999; 15:
407-425.
Varner J, Ramkrishna D. Metabolic engineering from a cybernetic
perspective: I. Qualitative investigation of nodal architectures and
their response to genetic perturbations. Biotechnol Prog 1999; 15:
426-438.
Namjoshi A, Ramkrishna D. A cybernetic modeling framework for
analysis of metabolic systems. Comp Chem Eng 2005; 29: 487498.
Goodwin BC. Oscillatory organization in cells, a dynamic theory of
cellular control processes. Academic Press, New York, 1963.
Savageau MA. Biochemical systems analysis. I. Some mathematical properties of the rate laws for the component enzymatic reactions. J Theor Biol 1969; 25: 365-369.
Savageau MA. Biochemical systems analysis. II. The steadystate
solutions for an n-pool system using a power-law approximation. J
Theor Biol 1969; 25: 370-379.
Savageau MA. Biochemical systems analysis. III. Dynamic solutions using a power-law approximation. J Theor Biol 1970; 26:
215-226.
Hatzimanikatis V, Bailey JE. MCA has more to say. J Theor Biol
1996; 182: 233-242.
Stephanopoulos G, Aristidou AA, Nielsen J. Metabolic engineering, principles and methodologies. Academic Press, San Diego,
1998.
Edwards JS, Ramakrishna R, Schilling CH, Palsson BO. Metabolic
flux balance analysis. In Metabolic Engineering, Lee SY and Papoutsakis ET, editors. Marcel Dekker, New York, 1999; pp. 13-57.
Kompala DS. Bacterial growth on multiple substrates. Experimental verification of cybernetic models. Ph. D. thesis. Purdue University, West Lafayette, IN, 1984.
Dhurjati P, Ramkrishna D, Flickinger MC, Tsao GT. A cybernetic
view of microbial growth: Modeling of cells as optimal strategists.
Biotech Bioeng 1985; 27: 1-9.
Kompala D, Jansen N, Tsao G, Ramkrishna D. Investigation of
bacterial growth on mixed substrates: Experimental evaluation of
cybernetic models. Biotech Bioeng 1986; 28: 1044-1055.
Holzhtter HG. The principle of flux minimization and its application to estimate stationary fluxes in metabolic networks. Eur J Biochem 2004; 271: 2905-2922.
Heijnen J. Approximative kinetic formats used in metabolic network modeling. Biotechnol Bioeng 2005; 91: 534-545.
Vaseghi S, Baumeisters A, Rizzi M, Reuss M. In vivo dynamics of
the pentose phosphate pathway in Saccharomyces cerevisiae. Metab Eng 1999; 1: 128-140.
Teusink B, Passarge J, Reijenga CA, et al. Can yeast glycolysis be
understood in terms of in vitro kinetics of the constituent enzymes?
Testing biochemistry. Eur J Biochem 2000; 267: 5313-5329.
Nielsen J, Jorgensen HS. Metabolic control analysis of the penicillin biosynthetic pathway in a high-yielding strain of Penicillium
chrysogenum. Biotechnol Prog 1995; 11: 299-305.
Cakir T, Patil KR, Onsan ZI, Ulgen KO, Kirdar B, Nielsen J. Integration of metabolome data with metabolic networks reveals reporter reactions. Mol Syst Biol 2006; 2: doi:10.1038/msb4100085.
Tomita M, Hashimoto K, Takahashi K, Shimizu TS, Matsuzaki Y,
Miyoshi F, Saito K, Tanida S, Yugi K, Venter JC, Hutchison CA
3rd. E-CELL: Software environment for whole-cell simulation.
Bioinformatics 1999; 15: 72-84.
Chong L, Ray LB. Whole-istic biology. Science 2002; 295: 1661.
Ideker T, Galitski T, Hood LA. New approach to decoding life:
Systems biology. Annu Rev Genomics Hum Genet 2001; 2: 343372.
Kitano H. Systems biology: A brief overview. Science 2002; 295:
1662-1664.
Oltvai ZN, Barabasi AL. Systems biology. Lifes complexity
pyramid. Science 2002; 298: 763-764.
Henry CM. Systems biology. Chem Eng News 2003; 81: 45-55.
Hood L. Systems biology: Integrating technology, biology and
computation. Mech Ageing Dev 2003; 124: 9-16.
Ideker T. Systems biology 101-What you need to know. Nature
Biotechnol 2004; 22: 473-475.
Aderem A. Systems biology: Its practice and challenges. Cell 2005;
121: 511-513.
Kirschner MW. The meaning of systems biology. Cell 2005; 121:
503-504.

Emerging Roles for Metabolic Engineering


[69]
[70]

[71]
[72]
[73]
[74]
[75]
[76]
[77]
[78]

[79]
[80]

[81]
[82]

[83]

[84]
[85]

[86]
[87]

[88]

[89]
[90]

[91]
[92]

[93]
[94]

[95]

Liu ET. Systems biology, integrative biology, predictive biology.


Cell 2005; 121: 505-506.
Westerhoff HV, Boogerd FC, Bruggeman FJ, Richardson RC,
Stephan A, Westerhoff HV. Systems biology in action. Curr Opin
Biotechnol 2005; 16: 326-328.
Williamson MP. Systems biology: Will it work? Biochem Soc
Trans 2005; 33: 503-506.
Pavenstadt H, Kriz W, Kretzler M. Cell biology of the glomerular
podocyte. Physiol Rev 2003; 83: 253-307.
Eddy AA, Schnaper HW. The nephrotic syndrome: From the simple to the complex. Semin Nephrol 1998; 18: 304-316.
Kriz W, Lemley KV. The role of the podocyte in glomerulosclerosis. Curr Opin Neph Hyperten 1999; 8: 489-497.
Somlo S, Mundel P. Getting a foothold in nephrotic syndrome. Nat
Genet 2000; 24: 333-335.
Endlich K, Kriz W, Witzgall R. Update in podocyte biology. Curr
Opin Nephrol Hypertens 2001; 10: 331-340.
Shirato I, Sakai T, Kimura K, Tomino Y, Kriz W. Cytoskeletal
changes in podocytes associated with foot process effacement in
Masugi nephritis. Am J Pathol 1996; 148: 1283-1296.
Smoyer WE, Mundel P. Regulation of podocyte structure during
the development of nephrotic syndrome. J Mol Med 1998; 76: 172183.
Reiser J, Von Gersdorff G, Simons M, et al. Novel concepts in
understanding and management of glomerular proteinuria. Nephrol
Dial Transplant 2002; 17: 951-955.
Farquhar MG, Vernier RL, Good RA. An electron microscope
study of the glomerulus in nephrosis, glomerulonephritis, and lupus
erythematosus. J Exp Med 1957; 106: 649-660.
Seiler MW, Venkatachalam MA, Cotran RS. Glomerular epithelium: Structural alterations induced by polycations. Science 1975;
189: 390-393.
Seiler MR, Rennke HG, Venkatachalam MA, Cotran RS. Pathogenesis of polycation-induced alteration (fusion) of glomerular epithelium. Lab Invest 1977; 36: 48-61.
Orlando RA, Takeda T, Zak B, et al. The glomerular epithelial cell
anti-adhesin podocalyxin associates with the actin cytoskeleton
through interactions with ezrin. J Am Soc Nephrol 2001; 12: 15891598.
Takeda T, McQuistan T, Orlando RA, Farquhar MG. Loss of
glomerular foot processes is associated with uncoupling of podocalyxin from the actin cytoskeleton. J Clin Invest 2001; 108: 289-301.
Kawachi H, Kurihara H, Topham PS, et al. Slit diaphragm-reactive
nephritogenic MAb 5-1-6 alters expression of ZO-1 in rat podocytes. Am J Physiol Renal Physiol 1997; 273: F984-F993.
Fujigaki Y, Nagase M, Hidaka S, et al. Altered anionic GBM components in monoclonal antibody against slit diaphragm-injected
proteinuric rats. Kidney Int 1998; 54: 1491-1500.
Wharram BL, Goyal M, Gillespie PJ, et al. Altered podocyte structure in GLEPP1 (Ptpro)-deficient mice associated with hypertension and low glomerular filtration rate. J Clin Invest 2000; 106:
1281-1290.
Smoyer WE, Mundel P, Gupta A, Welsh MJ. Podocyte alphaactinin induction precedes foot process effacement in experimental
nephrotic syndrome. Am J Physiol 1997; 273: F150-157.
Kaplan J, Pollak MR. Familial focal segmental glomerulosclerosis.
Curr Opin Nephrol Hypertens 2001; 10: 183-187.
Noakes PG, Miner JH, Gautam M, Cunningham JM, Sanes JR,
Merlie JP. The renal glomerulus of mice lacking s-laminin/laminin
beta-2: Nephrosis despite molecular compensation by laminin beta1. Nat Genet 1995; 10: 400-406.
Kreidberg JA, Donovan MJ, Goldstein SL, et al. Alpha-3 beta-1
integrin has a crucial role in kidney and lung organogenesis. Development 1996; 122: 3537-3547.
Miner JH, Sanes JR. Molecular and functional defects in kidneys of
mice lacking collagen alpha-3(IV): Implications for Alport syndrome. J Cell Biol 1996; 135: 1403-1413.
Miner JH, Li C. Defective glomerulogenesis in the absence of
laminin alpha-5 demonstrates a developmental role for the kidney
glomerular basement membrane. Dev Biol 2000; 217: 278-289.
Raats CJ, Van Den Born J, Berden JH. Glomerular heparan sulfate
alterations: Mechanisms and relevance for proteinuria. Kidney Int
2000; 57: 385-400.
Regele HM, Fillipovic E, Langer B, et al. Glomerular expression of
dystroglycans is reduced in minimal change nephrosis but not in

Current Chemical Biology, 2008, Vol. 2, No. 1

[96]

[97]

[98]
[99]

[100]

[101]

[102]
[103]

[104]
[105]

[106]

[107]

[108]
[109]

[110]
[111]

[112]

[113]
[114]

[115]
[116]
[117]

[118]

79

focal segmental glomerulosclerosis. J Am Soc Nephrol 2000; 11:


403-412.
Kretzler M, Teixeira VP, Unschuld PG, et al. Integrin-linked
kinase as a candidate downstream effector in proteinuria. FASEB J
2001; 15: 1843-1845.
Morello R, Zhou G, Dreyer SD, et al. Regulation of glomerular
basement membrane collagen expression by LMX1B contributes to
renal disease in nail patella syndrome. Nat Genet 2001; 27: 205208.
Kestila M, Lenkkeri U, Mannikko M, Lamerdin J, et al. Positionally cloned gene for a novel glomerular protein -nephrin- is mutated in congenital nephrotic syndrome. Mol Cell 1998; 1: 575-582.
Lenkkeri U, Mannikko M, McCready P, et al. Structure of the gene
for congenital nephrotic syndrome of the Finnish type (NPHS1)
and characterization of mutations. Am J Hum Genet 1999; 64: 5161.
Liu G, Kaw B, Kurfis J, Rahmanuddin S, Kanwar YS, Chugh SS.
Neph1 and nephrin interaction in the slit diaphragm is an important
determinant of glomerular permeability. J Clin Invest 2003; 112:
209-221.
Reiser J, Mundel P. Danger signaling by glomerular podocytes
defines a novel function of inducible B7-1 in the pathogenesis of
nephrotic syndrome. J Am Soc Nephrol 2004; 15: 2246-2248.
Reiser J, Oh J, Shirato I, et al. Podocyte migration during nephrotic
syndrome requires a coordinated interplay between cathepsin L and
alpha-3 integrin. J Biol Chem 2004; 279: 34827-34832.
Reiser J, Von Gersdorff G, Loos M, et al. Induction of B7-1 in
podocytes is associated with nephrotic syndrome. J Clin Invest
2004; 113: 1390-1397.
Reiser J, Pixley FJ, Hug A, et al. Regulation of mouse podocyte
process dynamics by protein tyrosine phosphatases. Kidney Int
2000; 57: 2035-2042.
Wiechert W, Mllney M, Petersen S, de Graaf AA. Significant
progress has been made in using existing metabolic databases to estimate metabolic fluxes. A universal framework for 13C metabolic
flux analysis. Metab Eng 2001; 3: 265-283.
Klapa MI, Park SM, Sinskey AJ, Stephanopoulos G. Metabolite
and isotopomer balancing in the analysis of metabolic cycles: I.
Theory. Biotechnol Bioeng 1999; 62: 375-391.
Bonarius HPJ, Hatzimanikatis V, Meesters KPH, de Gooijer CD,
Schmid G, Tramper J. Metabolic flux analysis of hybridoma cells
in different culture media using mass balances. Biotechnol Bioeng
1996; 50: 299-318.
Xie L, Wang DC. Material balance studies on animal cell metabolism using a stoichiometrically based reaction network. Biotechnol
Bioeng 1996; 52: 579-590.
Follstad BD, Balcarcel R, Wang DIC, Stephanopoulos G. Metabolic flux analysis of hybridoma continuous culture steady-state
multiplicity. Biotechnol Bioeng 1999; 63: 675-683.
Balcarcel RR, Clark LM. Metabolic screening of mammalian cell
cultures using well-plates. Biotechnol Prog 2003; 19: 98-108.
Nyberg GB, Balcarcel RP, Follstad BD, Stephanopoulos G, Wang
DIC. Metabolism of peptide amino acids by chinese hamster ovary
cells grown in a complex medium. Biotechnol Bioeng 1999; 62:
324-335.
Nadeau I, Sabatie J, Koehl M, Perrier M, Kamen A. Human 293
cell metabolism in low glutamine-supplied culture: Interpretation
of metabolic changes through metabolic flux analysis. Metab Eng
2000; 2: 277-292.
Nadeau I, Jacob D, Perrier M, Kamen A. 293SF metabolic flux
analysis during cell growth and infection with an adenoviral vector.
Biotechnol Prog 2000; 16: 872-884.
Chan C, Berthiaume F, Lee K, Yarmush ML. Metabolic flux analysis of cultured hepatocytes exposed to plasma. Biotechnol Bioeng
2003; 81: 33-49.
Arai K, Lee K, Berthiaume F, Tompkins RG, Yarmush ML. Intrahepatic amino acid and glucose metabolism in a D-galactosamine
induced rat liver failure model. Hepatology 2001; 34: 360-371.
Kahn P. From genome to proteome: Looking at a cells proteins.
Science 1995; 270: 369-370.
Wilkins MR, Sanchez JC, Gooley AA, et al. Progress with proteome projects: Why all proteins expressed by a genome should be
identified and how to do it. Biotechnol Genet Eng Rev 1996; 13:
19-50.
Liebler DC. Introduction to proteomics: Tools for the new biology.
Humana Press, Totowa, NJ, 2001.

80 Current Chemical Biology, 2008, Vol. 2, No. 1


[119]

[120]

[121]
[122]

[123]
[124]
[125]
[126]
[127]

[128]
[129]

[130]
[131]

[132]

[133]
[134]

[135]

[136]
[137]

[138]

[139]
[140]

[141]
[142]

Palmer-Toy DE, Kuzdzal S, Chan DW. Proteomic approaches to


tumor marker discovery. In Tumor markers: Physiology, pathobiology, Technology and clinical applications, Diamandis EP,
Fritsche HA, Lilja H, Chan DW, Schwartz M (eds.), AACC Press,
Washington DC, 2002; pp. 391-400.
Yan JX, Devenish AT, Wait R, Stone T, Lewis S, Fowler S. Fluorescence two-dimensional difference gel electrophoresis and mass
spectrometry based proteomic analysis of Escherichia coli. Proteomics 2002; 2: 1682-1698.
Lilley KS, Dupree P. Methods of quantitative proteomics and their
application to plant organelle characterization. J Exp Bot 2006; 57:
1493-1499.
Wolters DA, Washburn MP, Yates JR 3rd. An automated multidimensional protein identification technology for shotgun proteomics. Anal Chem 2001; 73: 5683-5690.
Ransom, RF. Podocyte proteomics. Contrib Nephrol 2004; 141:
189-211.
Ransom RF, Vega-Warner V, Smoyer WE, Klein J. Differential
proteomic analysis of proteins induced by glucocorticoids in cultured murine podocytes. Kidney Int 2005; 67: 1275-1285.
Dreger M. Proteome analysis at the level of subcellular structures.
Eur J Biochem 2003; 270: 589-599.
Chen P, Harcum SW. Effects of amino acid additions on ammonium stressed CHO cells. J Biotechnol 2005; 117: 277-286.
Linz M, Zeng AP, Wagner R, Deckwer WD. Stoichiometry, kinetics and regulation of glucose and amino acid metabolism of a recombinant BHK cell line in batch and continuous cultures. Biotechnol Prog 1997; 13: 453-463.
Schneider M, Marison IW, von Stockar U. The importance of ammonia in mammalian cell culture. J Biotechnol 1996; 46: 161-185.
Zeilke HR, Ozand PT, Tildon JT, Sevdalian DA, Cornblath M.
Reciprocal regulation of glucose and glutamine utilization by cultured human diploid fibroblasts. J Cell Physiol 1978; 95: 41-48.
Glacken MW, Huang C, Sinskey AJ. Mathematical descriptions of
hybridoma culture kinetics: III. Simulation of fed-batch bioreactors. J Biotechnol 1989; 10: 39-66.
Fitzpatrick L, Jenkins HA, Butler M. Glucose and glutamine metabolism of a murine B-lymphocyte hybridoma grown in batch culture. Appl Biochem Biotechnol 1993; 43: 93-116.
Elias CB, Carpentier E, Durocher Y, Bisson L, Wagner R, Kamen
A. Improving glucose and glutamine metabolism of human HEK
293 and Trichoplusia ni insect cells engineered to express a cytosolic pyruvate carboxylase enzyme. Biotechnol Prog 2003; 19: 90-97.
Neermann J, Wagner R. Comparative analysis of glucose and glutamine metabolism in transformed mammalian cell lines, insect and
primary liver cells. J Cell Phys 1996; 166: 152-169.
Calik P, Calik G, Takac S, Ozdamar TH. Metabolic flux analysis
for serine alkaline protease fermentation by Bacillus licheniformis
in a defined medium: Effects of the oxygen transfer rate. Biotechnol Bioeng 1999; 64: 151-167.
Dauner M, Storni T, Sauer U. Bacillus subtilis metabolism and
energetics in carbon-limited and excess-carbon chemostat culture. J
Bacteriol 2001; 183: 7308-7317.
Goel A, Ferrance J, Jeong J, Ataa MM. Analysis of metabolic
fluxes in batch and continuous cultures of Bacillus subtilis. Biotechnol Bioeng 1993; 42: 686-696.
Sauer U, Hatzimanikatis V, Hohmann HP, Manneberg M, van
Loon AP, Bailey JE. Physiology and metabolic fluxes of wild-type
and riboflavin-producing Bacillus subtilis. Appl Environ Microbiol
1996; 62: 3687-3696.
Sauer U, Hatzimanikatis V, Bailey JE, Hochuli M, Szyperski T,
Wuthrich K. Metabolic fluxes in riboflavin-producing Bacillus subtillis. Nat Biotechnol 1997; 15: 448-452.
Dauner M, Bailey JE, Sauer U. Metabolic flux analysis with a
comprehensive isotopomer model in Bacillus subtilis. Biotechnol
Bioeng 2001; 76: 144-156.
Zhu Y, Chen X, Chen T, Zhao X. Enhancement of riboflavin production by overexpression of acetolactate synthase in a pta mutant
of Bacillus subtilis. FEMS Microbiol Lett 2007; 266: 224-230.
Reardon KF, Scheper TH, Bailey JE. Metabolic pathway rates and
culture fluorescence in batch fermentations of Clostridium acetobutylicum. Biotechnol Prog 1987; 3: 153-167.
Desai RP, Harris LM, Welker NE, Papoutsakis ET. Metabolic flux
analysis elucidates the importance of acid-formation pathways in
regulating solvent production by Clostridium acetobutylicum. Metab Eng 1999; 1: 206-213.

Altintas et al.
[143]

[144]

[145]

[146]

[147]
[148]
[149]

[150]

[151]
[152]

[153]
[154]

[155]
[156]

[157]

[158]

[159]
[160]

[161]
[162]

[163]
[164]
[165]
[166]

Desai RP, Nielsen LK, Papoutsakis ET. Stoichiometric modeling of


Clostridium acetobutylicum fermentations with non-linear constraints. J Biotechnol 1999; 71: 191-205.
Harris LM, Blank L, Desai RP, Welker NE, Papoutsakis ET. Fermentation characterization and flux analysis of recombinant strains
of Clostridium acetobutylicum with an inactivated solR gene. J Ind
Microbiol Biotechnol 2001; 27: 322-328.
Desvaux M, Petitdemange H. Flux analysis of the metabolism of
Clostridium cellulolyticum grown in cellulose-fed continuous culture on a chemically defined medium under ammonium-limited
conditions. Appl Environ Microbiol 2001; 67: 3846-3851.
Desvaux M, Guedon E, Petitdemange H. Metabolic flux in cellulose batch and cellulose-fed continuous cultures of Clostridium cellulolyticum in response to acidic environment. Microbiology 2001;
147: 1461-1471.
Zupke C, Stephanopoulos G. Modeling of isotope distributions and
intracellular fluxes in metabolic networks using atom mapping matrices. Biotechnol Prog 1994; 10: 489-498.
Kimura E. Metabolic engineering of glutamate production. Adv
Biochem Eng Biotechnol 2003; 79: 37-57.
Shirai T, Matsuzaki K, Kuzumoto M, et al. Precise metabolic flux
analysis of Coryneform bacteria by gas chromatographymass
spectrometry and verification by nuclear magnetic resonance. J Biosci Bioeng 2006; 102: 413-424.
Vallino JJ, Stephanopoulos G. Metabolic flux distributions in
Corynebacterium glutamicum during growth and lysine overproduction. Biotechnol Bioeng 1993; 41: 633-646.
Vallino JJ, Stephanopoulos G. Carbon flux distributions at the
glucose-6-phosphate branch point in Corynebacterium glutamicum
during lysine overproduction. Biotechnol Prog 1994; 10: 327-334.
Vallino JJ, Stephanopoulos G. Carbon flux distributions at the
pyruvate branch point in C. glutamicum during lysine overproduction. Biotechnol Prog 1994; 10: 327-334.
Vallino JJ, Stephanopoulos G. Metabolic flux distributions in
Corynebacterium glutamicum during growth and lysine overproduction. Biotechnol Bioeng 2000; 67: 872-885.
Marx A, de Graaf AA, Wiechert W, Eggeling L, Sahm H. Determination of the fluxes in the central metabolism of Corynebacterium
glutamicum. Biotechnol Bioeng 1996; 49: 111-129.
Wittmann C, Heinzle E. Modeling and experimental design for
metabolic flux analysis of lysine-producing Corynebacteria by
mass spectrometry. Metab Eng 2001; 3: 173-191.
Drysch A, El Massaoudi M, Mack C, Takors R, de Graaf AA,
Sahm H. Production process monitoring by serial mapping of microbial carbon flux distributions using a novel sensor reactor approach: II-13C-labeling-based metabolic flux analysis and L-lysine
production. Metab Eng 2003; 5: 96-107.
Yang J, Wongsa S, Kadirkamanathan V, Billings SA, Wright PC.
Metabolic flux estimation - A self-adaptive evolutionary algorithm
with singular value decomposition. IEEE/ACM Trans Comp Biol
Bioinformatics 2007; 4: 126-138.
Kiefer P, Heinzle E, Zelder O, Wittmann C. Comparative metabolic flux analysis of lysine-producing Corynebacterium glutamicum cultured on glucose or fructose. Appl Environ Microbiol
2004; 70: 229-239.
Wittmann C, Kiefer P, Zelder O. Metabolic fluxes in Corynebacterium glutamicum during lysine production with sucrose as carbon
source. Appl Environ Microbiol 2004; 70: 7277-7287.
Park SM, Sinskey AJ, Stephanopoulos G. Metabolic and physiological studies of Corynebacterium glutamicum mutants. Biotechnol Bioeng 1997; 55: 864-879.
Ohnishi J, Katahira R, Mitsuhashi S, Kakita S, Ikeda M. A novel
gnd mutation leading to increased L-lysine production in Corynebacterium glutamicum. FEMS Microbiol Lett 2005; 242: 265-274.
Pons A, Dussap CG, Pequignot C, Gros JB. Metabolic flux distribution in Cornybacterium melassecola ATCC 17965 for various
carbon sources. Biotechnol Bioeng 1996; 51: 177-189.
Majewski RA, Domach MM. Simple constrained optimization
view of acetate overflow in Escherichia coli. Biotechnol Bioeng
1990; 35: 732-738.
Delgado J, Liao JC. Inverse flux analysis for reduction of acetate
excretion in Escherichia coli. Biotechnol Prog 1997; 13: 361-367.
Varma A, Palsson BO. Metabolic capabilities of Escherichia coli.
II. Optimal growth patterns. J Theor Biol 1993; 165: 503-522.
Aristidou AA, San KY, Bennett GN. Metabolic flux analysis of
Escherichia coli expressing the Bacillus subtilis acetolactate syn-

Emerging Roles for Metabolic Engineering

[167]

[168]

[169]
[170]

[171]

[172]

[173]

[174]

[175]
[176]

[177]
[178]

[179]

[180]
[181]

[182]
[183]

[184]

[185]
[186]

[187]

thase in batch and continuous cultures. Biotechnol Bioeng 1999;


63: 737-749.
Yang YT, Aristidou AA, San KY, Bennett GN. Metabolic flux
analysis of Escherichia coli deficient in the acetate production
pathway and expressing the Bacillus subtilis acetolactate synthase.
Metab Eng 1999; 1: 26-34.
Edwards JS, Palsson BO. Metabolic flux balance analysis and the
in silico analysis of Escherichia coli K-12 gene deletions. BMC
Bioinformatics 2000; 1: 1-10.
Burgard AP, Maranas CD. Probing the performance limits of the
Escherichia coli metabolic network subject to gene additions or deletions. Biotechnol Bioeng 2001; 74: 364-375.
Emmerling M, Dauner M, Ponti A, et al. Metabolic flux responses
to pyruvate kinase knockout in Escherichia coli. J Bacteriol 2002;
184: 152-164.
Hua Q, Yang C, Baba T, Mori H, Shimizu K. Responses of the
central metabolism in Escherichia coli to phosphoglucose
isomerase and glucose-6-phosphate dehydrogenase knockouts. J
Bacteriol 2003; 185: 7053-7067.
Yang C, Hua Q, Baba T, Mori H, Shimizu K. Analysis of Escherichia coli anaplerotic metabolism and its regulation mechanisms from the metabolic responses to altered dilution rates and
phosphoenolpyruvate carboxykinase knockout. Biotechnol Bioeng
2003; 84: 129-144.
Al Zaid Siddiquee K, Arauzo-Bravo MJ, Shimizu K. Metabolic
flux analysis of pykF gene knockout Escherichia coli based on 13Clabeling experiments together with measurements of enzyme activities and intracellular metabolite concentrations. Appl Microbiol
Biotechnol 2004; 63: 407-417.
Peng L, Arauzo-Bravo MJ, Shimizu K. Metabolic flux analysis for
a ppc mutant Escherichia coli based on 13C-labelling experiments
together with enzyme activity assays and intracellular metabolite
measurements. FEMS Microbiol Lett 2004; 235: 17-23.
Zhao J, Baba T, Mori H, Shimizu K. Effect of zwf gene knockout
on the metabolism of Escherichia coli grown on glucose or acetate.
Metab Eng 2004; 6: 164-174.
Diaz-Ricci JC, Tsu M, Bailey JE. Influence of expression of the pet
operon on intracellular metabolic fluxes of Escherichia coli. Biotechnol Bioeng 1992; 39: 59-65.
Varma A, Boesch BW, Palsson BO. Stoichiometric interpretation
of Escherichia coli glucose catabolism under various oxygenation
rates. Appl Environ Microb 1993; 59: 2465-2473.
Varma A, Palsson BO. Stoichiometric flux balance models quantitatively predict growth and metabolic byproduct secretion in wild
type Escherichia coli W3110. Appl Environ Microbiol 1994; 60:
3724-3731.
Varma A, Palsson BO. Parametric sensitivity of stoichiometric flux
balance models applied to wild type Escherichia coli metabolism.
Biotechnol Bioeng 1995; 45: 69-79.
Holms WH. Flux analysis and control of the central metabolic
pathways in Escherichia coli. FEMS Microbiol Rev 1996; 19: 85116.
Schilling CH, Edwards JS, Letscher D, Palsson BO. Combining
pathway analysis with flux balance analysis for the comprehensive
study of metabolic systems. Biotechnol Bioeng 2000; 71: 286-306.
Keasling JD, Van Dien SJ, Pramanik J. Engineering polyphosphate
metabolism in Escherichia coli: Implications for bioremediation of
inorganic contaminants. Biotechnol Bioeng 1998; 58: 231-239.
Hong SH, Lee SY. Metabolic flux analysis for succinic acid production by recombinant Escherichia coli with amplified malic enzyme activity. Biotechnol Bioeng 2001; 74: 89-95.
Lee SY, Hong SH, Moon SY. In silico metabolic pathway analysis
and design: Succinic acid production by metabolically engineered
Escherichia coli as an example. Genome Inform 2002; 13: 214223.
Sanchez AM, Bennett GN, San KY. Batch culture characterization
and metabolic flux analysis of succinate-producing Escherichia
coli strains. Metab Eng 2006; 8: 209-226.
Bai DM, Zhao XM, Li XG, Xu SM. Strain improvement and metabolic flux analysis in the wild-type and a mutant Lactobacillus lactis strain for L(+)-lactic acid production. Biotechnol Bioeng 2004;
88: 681-689.
Naeimpoor F, Mavituna F. Metabolic flux analysis in Streptomyces
coelicolor under various nutrient limitations. Metab Eng 2000; 2:
140-148.

Current Chemical Biology, 2008, Vol. 2, No. 1


[188]

[189]
[190]

[191]

[192]

[193]

[194]
[195]

[196]
[197]

[198]
[199]

[200]

[201]
[202]

[203]
[204]

[205]
[206]

[207]

[208]

[209]

[210]

81

Kim HB, Smith CP, Micklefield J, Mavituna F. Metabolic flux


analysis for calcium dependent antibiotic (CDA) production in
Streptomyces coelicolor. Metab Eng 2004; 6: 313-325.
Daae EB, Ison AP. Classification and sensitivity analysis of a proposed primary metabolic reaction network for Streptomyces
lividans. Metab Eng 1999; 1: 153-165.
Avignone Rossa C, White J, Kuiper A, Postma PW, Bibb M, Teixeira de Mattos MJ. Carbon flux distribution in antibioticproducing chemostat cultures of Streptomyces lividans. Metab Eng
2002; 4: 138-150.
De Graaf AA, Striegel K, Wittig RM, et al. Metabolic state of
Zymomonas mobilis in glucose-, fructose-, and xylose-fed continuous cultures as analysed by 13C- and 31P-NMR spectroscopy. Arch
Microbiol 1999; 171: 371-385.
Kumar S, Punekar NS, SatyaNarayan V, Venkatesh KV. Metabolic
fate of glutamate and evaluation of flux through the 4aminobutyrate (GABA) shunt in Aspergillus niger. Biotechnol Bioeng 2000; 67: 575-584.
Schmidt K, Marx A, de Graaf AA, et al. 13C tracer experiments and
metabolite balancing for metabolic flux analysis: Comparing two
approaches. Biotechnol Bioeng 1998; 58: 254-257.
Granstrom TB, Aristidou AA, Jokela J, Leisola M. Growth characteristics and metabolic flux analysis of Candida milleri. Biotechnol
Bioeng 2000; 70: 197-207.
Cao Z, Gao H, Liu M, Jiao P. Engineering the acetyl-CoA transportation system of Candida tropicalis enhances the production of dicarboxylic acid. Biotechnol J 2006; 1: 68-74.
Granstrom T, Aristidou AA, Leisola M. Metabolic flux analysis of
Candida tropicalis growing on xylose in an oxygen-limited chemostat. Metab Eng 2002; 4: 248-256.
Jorgensen H, Nielsen J, Villadsen J, Mollgaard H. Metabolic flux
distributions in Penicillium chrysogenum during fed-batch cultivations. Biotechnol Bioeng 1995; 46: 117-131.
Henriksen CM, Christensen LH, Nielsen J, Villadsen J. Growth
energetics and metabolic fluxes in continuous cultures of Penicillium chrysogenum. J Biotechnol 1996; 45: 149-164.
Christensen B, Nielsen J. Metabolic network analysis of Penicillium chrysogenum using 13C-labeled glucose. Biotechnol Bioeng
2000; 68: 652-659.
van Winden WA, van Gulik WM, Schipper D, et al. Metabolic flux
and metabolic network analysis of Penicillium chrysogenum using
2D [13C, 1H] COSY NMR measurements and cumulative bondomer
simulation. Biotechnol Bioeng 2003; 83: 75-92.
Van Gulik WM, Heijnen JJ. A metabolic network stoichiometry
analysis of microbial growth and product formation. Biotechnol
Bioeng 1995; 48: 681-698.
Nissen TL, Schulze U, Nielsen J, Villadsen J. Flux distributions in
anaerobic, glucose-limited continuous cultures of Saccharomyces
cerevisiae. Microbiology 1997; 143: 203-218.
Franzen CJ. Metabolic flux analysis of RQ-controlled microaerobic
ethanol production by Saccharomyces cerevisiae. Yeast 2003; 20:
117-132.
Hjersted JL, Henson MA. Optimization of fed-batch Saccharomyces cerevisiae fermentation using dynamic flux balance models.
Biotechnol Prog 2006; 22: 1239-1248.
Hjersted JL, Henson MA, Mahadevan R. Genome-scale analysis of
Saccharomyces cerevisiae metabolism and ethanol production in
fed-batch culture. Biotechnol Bioeng 2007; 97: 1190-1204.
Tai SL, Daran-Lapujade P, Luttik MAH, et al. Control of the glycolytic flux in Saccharomyces cerevisiae grown at low temperature: A multi-level analysis in anaerobic chemostat cultures. J Biol
Chem 2007; 282: 10243-10251.
Cortassa S, Aon JC, Aon MA. Fluxes of carbon, phosphorylation,
and redox intermediates during growth of Saccharomyces cerevisiae on different carbon sources. Biotechnol Bioeng 1995; 47:
193-208.
Lee TH, Kim MY, Ryu YW, Seo JH. Estimation of theoretical
yield for ethanol production from D-xylose by recombinant Saccharomyces cerevisiae using metabolic pathway synthesis algorithm. J Microbiol Biotechnol 2001; 11: 384-388.
Wahlbom CF, Eliasson A, Hahn-Hagerdal B. Intracellular fluxes in
a recombinant xylose-utilizing Saccharomyces cerevisiae cultivated anaerobically at different dilution rates and feed concentrations. Biotechnol Bioeng 2001; 72: 289-296.
Wahlbom CF, Hahn-Hagerdal B. Furfural, 5-hydroxymethyl furfural, and acetoin act as external electron acceptors during anaero-

82 Current Chemical Biology, 2008, Vol. 2, No. 1

[211]

[212]

[213]

[214]
[215]

[216]

[217]
[218]

[219]

[220]

[221]

[222]
[223]

[224]
[225]

[226]
[227]

[228]

[229]

Altintas et al.

bic fermentation of xylose in recombinant Saccharomyces cerevisiae. Biotechnol Bioeng 2002; 78: 172-178.
Pitkanen JP, Aristidou A, Salusjarvi L, Ruohonen L, Penttila M.
Metabolic flux analysis of xylose metabolism in recombinant Saccharomyces cerevisiae using continuous culture. Metab Eng 2003;
5: 16-31.
Sonderegger M, Jeppsson M, Hahn-Hagerdal B, Sauer U. Molecular basis for anaerobic growth of Saccharomyces cerevisiae on xylose, investigated by global gene expression and metabolic flux
analysis. Appl Environ Microbiol 2004; 70: 2307-2317.
Grotkjaer T, Christakopoulos P, Nielsen J, Olsson L. Comparative
metabolic network analysis of two xylose fermenting recombinant
Saccharomyces cerevisiae strains. Metab Eng 2005; 7: 437-444.
Aon JC, Cortassa S. Involvement of nitrogen metabolism in the
triggering of ethanol fermentation in aerobic chemostat cultures of
Saccharomyces cerevisiae. Metab Eng 2001; 3: 250-264.
Cakir T, Arga KY, Altintas MM, Ulgen K. Flux analysis of the
recombinant Saccharomyces cerevisiae YPB-G utilizing starch for
optimum ethanol production. Process Biochem 2004; 39: 20972108.
Kleijn RJ, Geertman JMA, Nfor BK, Ras C, Schipper D, Pronk JT,
Heijnen JJ, van Maris AJA, van Winden WA. Metabolic flux
analysis of a glycerol-overproducing Saccharomyces cerevisiae
strain based on GC-MS, LC-MS and NMR-derived 13C-labelling
data. FEMS Yeast Res 2007; 7: 216-231.
Jin S, Ye K, Shimizu K. Metabolic flux distributions in recombinant Saccharomyces cerevisiae during foreign protein production. J
Biotechnol 1997; 54: 161-174.
Moller K, Christensen B, Forster J, Pikur J, Nielsen J, Olsson L.
Aerobic glucose metabolism of Saccharomyces kluyveri: Growth,
metabolite production, and quantification of metabolic fluxes. Biotechnol Bioeng 2002; 77: 186-193.
Cruz HJ, Moreira JL, Carrondo MJT. Metabolic shifts by nutrient
manipulation in continuous cultures of BHK cells. Biotechnol Bioeng 1999; 66: 104-113.
Nyberg GB, Balcarcel RR, Follstad BD, Stephanopoulos G, Wang
DI. Metabolic effects on recombinant interferon-gamma glycosylation in continuous culture of Chinese hamster ovary cells. Biotechnol Bioeng 1999; 62: 336-347.
Altamirano C, Illanes A, Casablancas A, Gamez X, Cairo JJ, Godia
C. Analysis of CHO cells metabolic redistribution in a glutamatebased defined medium in continuous culture. Biotechnol Prog
2001; 17: 1032-1041.
Altamirano C, Illanes A, Becerra S, Cairo JJ, Godia F. Considerations on the lactate consumption by CHO cells in the presence of
galactose. J Biotechnol 2006; 125: 547-556.
Henry O, Perrier M, Kamen A. Metabolic flux analysis of HEK293 cells in perfusion cultures for the production of adenoviral vectors. Metab Eng 2005; 7: 467-476.
Xie L, Wang D. Energy metabolism and ATP balance in animal
cell cultivation using a stoichiometrically based reaction network.
Biotechnol Bioeng 1996; 52: 591-601.
Zupke C, Sinskey AJ, Stephanopoulos G. Intracellular flux analysis
applied to the effect of dissolved oxygen on hybridomas. Appl Microbiol Biotechnol 1995; 44: 27-36.
Savinell JM, Palsson BO. Network analysis of intermediary metabolism using linear optimization: II. Interpretation of hybridoma
cell metabolism. J Theor Biol 1992; 154: 455-473.
Zupke C, Stephanopoulos G. Intracellular flux analysis in hybridomas using mass balances and in vitro 13C NMR. Biotechnol Bioeng 1995; 45: 292-303.
Bonarius HPJ, Ozemre A, Timmerarends B, et al. Metabolic-flux
analysis of continuously cultured hybridoma cells using 13CO(2)
mass spectrometry in combination with 13C-lactate nuclear magnetic resonance spectroscopy and metabolite balancing. Biotechnol
Bioeng 2001; 74: 528-538.
Xie L, Wang D. Applications if improved stoichiometric model in
medium design and fed-batch cultivation of animal cells in bioreactor. Cytotechnology 1994; 15: 17-29.

Received: December 7, 2006

[230]

[231]
[232]

[233]
[234]

[235]
[236]
[237]

[238]

[239]

[240]
[241]

[242]
[243]

[244]
[245]

[246]

[247]
[248]

[249]
[250]

Xie L, Wang D. Stoichiometric analysis of animal cell growth and


its application in medium design. Biotechnol Bioeng 1994; 43:
1164-1174.
Xie L, Wang D. High cell density and high monoclonal antibody
production through medium design and rational control in a bioreactor. Biotechnol Bioeng 1996; 51: 725-729.
Xie L, Wang D. Integrated approaches to the design of media and
feeding strategies for fed-batch cultures of animal cells. Trends
Biotechnol 1997; 15: 109-113.
De Alwis DM, Dutton RL, Scharer J, Moo-Young M. Statistical
methods in media optimization for batch and fed-batch animal cell
culture. Bioprocess Biosyst Eng 2007; 30: 107-113.
Gambhir A, Korke R, Lee J, Fu PC, Europa A, Hu WS. Analysis of
cellular metabolism of hybridoma cells at distinct physiological
states. J Biosci Bioeng 2003; 95: 317-327.
Si Y, Yoon J, Lee K. Flux profile and modularity analysis of timedependent metabolic changes of de novo adipocyte formation. Am
J Physiol Endocrinol Metab 2007; 292: E1637-E1646.
Fell DA, Small JR. Fat synthesis in adipose tissue. An examination
of stoichiometric constraints. Biochem J 1986; 238: 781-786.
Chatziioannou A, Palaiologos G, Kolisisa FN. Metabolic flux
analysis as a tool for the elucidation of the metabolism of neurotransmitter glutamate. Metab Eng 2003; 5: 201-210.
Alsan S, Cakir T, Saybasili H, Akin A, Ulgen K. Modelling of
neuron-astrocyte coupling via stoichiometric modelling technique.
Poster presented at ESBES5, Stuttgart-Germany, 8-11 September
2004.
Vo TD, Greenberg HJ, Palsson BO. Reconstruction and functional
characterization of the human mitochondrial metabolic network
based on proteomic and biochemical data. J Biol Chem 2004; 279:
39532-39540.
Thiele I, Price ND, Vo TD, Palsson BO. Candidate metabolic network states in human mitochondria: Impact of diabetes, ischemia
and diet. J Biol Chem 2005; 280: 11683-11695.
Lee K, Berthiaume F, Stephanopoulos GN, Yarmush ML. Induction of a hypermetabolic state in cultured hepatocytes by glucagon
and H2O2. Metab Eng 2003; 5: 221-229.
Lee K, Berthiaume F, Stephanopoulos GN, Yarmush ML. Profiling
of dynamic changes in hypermetabolic livers. Biotechnol Bioeng
2003; 83: 400-415.
Nolan RP, Fenley AP, Lee K. Identification of distributed metabolic objectives in the hypermetabolic liver by flux and energy balance analysis. Metab Eng 2006; 8: 30-45.
Calik P, Akbay A. Mass flux balance-based model and metabolic
flux analysis for collagen synthesis in the fibrogenesis process of
human liver. Med Hypoth 2000; 55: 5-14.
Chan C, Berthiaume F, Lee K, Yarmush ML. Metabolic flux analysis of hepatocyte function in hormone- and amino acid supplemented plasma. Metab Eng 2003; 5: 1-15.
Yokoyama T, Banta S, Berthiaume F, Nagrath D, Tompkins RG,
Yarmush ML. Evolution of intrahepatic carbon, nitrogen and energy metabolism in a D-galactosamine-induced rat liver failure
model. Metab Eng 2005; 7: 88-103.
Yamaguchi Y, Yu YM, Zupke C, et al. Effect of burn injury on
glucose and nitrogen metabolism in the liver: Preliminary studies in
a perfused liver system. Surgery 1997; 121: 295-303.
Lee K, Berthiaume F, Stephanopoulos GN, Yarmush DM, Yarmush ML. Metabolic flux analysis of postburn hepatic hypermetabolism. Metab Eng 2000; 2: 312-327.
Banta S, Yokoyama T, Berthiaume F, Yarmush ML. Effects of
dehydroepiandrosterone administration on rat hepatic metabolism
following thermal injury. J Surg Res 2005; 127: 93-105.
Banta S, Yokoyama T, Berthiaume F, Yarmush ML. Quantitative
effects of thermal injury and insulin on the metabolism of the
skeletal muscle using the perfused rat hindquarter preparation. Biotechnol Bioeng 2004; 88: 613-629.

Revised: February 1, 2007

Accepted: March 12, 2007

Vous aimerez peut-être aussi