Linking high-resolution metabolic flux phenotypes and

advertisement
Linking high-resolution metabolic flux phenotypes and
transcriptional regulation in yeast modulated by the global
regulator Gcn4p
The MIT Faculty has made this article openly available. Please share
how this access benefits you. Your story matters.
Citation
Moxley, Joel F et al. “Linking high-resolution metabolic flux
phenotypes and transcriptional regulation in yeast modulated by
the global regulator Gcn4p.” Proceedings of the National
Academy of Sciences 106.16 (2009): 6477-6482.
As Published
http://dx.doi.org/10.1073/pnas.0811091106
Publisher
National Academy of Sciences
Version
Final published version
Accessed
Fri May 27 04:46:43 EDT 2016
Citable Link
http://hdl.handle.net/1721.1/50252
Terms of Use
Article is made available in accordance with the publisher's policy
and may be subject to US copyright law. Please refer to the
publisher's site for terms of use.
Detailed Terms
SPECIAL FEATURE
Linking high-resolution metabolic flux phenotypes
and transcriptional regulation in yeast modulated
by the global regulator Gcn4p
Joel F. Moxleya,1, Michael C. Jewettb,1, Maciek R. Antoniewicza,1,2, Silas G. Villas-Boasb,1,3, Hal Alpera,c,
Robert T. Wheelerd, Lily Tonga, Alan G. Hinnebusche, Trey Idekerd, Jens Nielsenb, and Gregory Stephanopoulosa,4
aDepartment of Chemical Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139; bCenter for Microbial Biotechnology, Technical
University of Denmark, BioCentrum-DTU, Building 223, DK-2800 Kongens Lyngby, Denmark; cWhitehead Institute for Biomedical Research, Massachusetts
Institute of Technology, 9 Cambridge Center, Cambridge, MA 02139; eLaboratory of Gene Regulation and Development, National Institute of Child Health
and Human Development, Bethesda, MD 20892; and dDepartment of Bioengineering, University of California, San Diego, CA 92093
Genome sequencing dramatically increased our ability to understand cellular response to perturbation. Integrating system-wide
measurements such as gene expression with networks of protein–
protein interactions and transcription factor binding revealed critical insights into cellular behavior. However, the potential of
systems biology approaches is limited by difficulties in integrating
metabolic measurements across the functional levels of the cell
despite their being most closely linked to cellular phenotype. To
address this limitation, we developed a model-based approach to
correlate mRNA and metabolic flux data that combines information
from both interaction network models and flux determination
models. We started by quantifying 5,764 mRNAs, 54 metabolites,
and 83 experimental 13C-based reaction fluxes in continuous cultures of yeast under stress in the absence or presence of global
regulator Gcn4p. Although mRNA expression alone did not directly
predict metabolic response, this correlation improved through
incorporating a network-based model of amino acid biosynthesis
(from r ⴝ 0.07 to 0.80 for mRNA-flux agreement). The model
provides evidence of general biological principles: rewiring of
metabolic flux (i.e., use of different reaction pathways) by transcriptional regulation and metabolite interaction density (i.e., level
of pairwise metabolite-protein interactions) as a key biosynthetic
control determinant. Furthermore, this model predicted flux rewiring in studies of follow-on transcriptional regulators that were
experimentally validated with additional 13C-based flux measurements. As a first step in linking metabolic control and genetic
regulatory networks, this model underscores the importance of
integrating diverse data types in large-scale cellular models. We
anticipate that an integrated approach focusing on metabolic
measurements will facilitate construction of more realistic models
of cellular regulation for understanding diseases and constructing
strains for industrial applications.
amino acid stress response 兩 fluxomics 兩 gcn4 兩 systems biology
M
etabolic fluxes are informative indicators of cellular physiology and in vivo homeostasis. Fluxes are the modelindependent rates of metabolite interconversion that emerge
through the interplay of genes, proteins, and metabolites at
multiple regulatory levels. Protein–protein interactions (1) and
transcription factor binding (2, 3) can be used to predict a
cellular behavior at the gene and protein level (4, 5). However,
complex metabolite-enzyme interactions (6), translational regulation (7), and posttranscriptional mechanisms prevent direct
linkage of transcriptional state and metabolic phenotype. Previous in silico attempts to address this gap included metabolic
network and regulatory on/off switch models for growth and
viability prediction (8, 9). In addition, experimental studies of
posttranscriptional control have measured mRNA expression in
conjunction with either 13C-based flux (10) or metabolite level
measurements (11). Such studies have proven invaluable for
www.pnas.org兾cgi兾doi兾10.1073兾pnas.0811091106
emphasizing the importance of metabolic phenotypes; however,
new approaches are needed which, by integrating these diverse
data types in biological models, can identify specific mechanisms
by which genetic regulatory circuits mediate metabolic flux
phenotype.
Toward this goal, we created a pathway model of amino acid
biosynthesis that includes genetic regulatory circuits and metabolite-enzyme interactions to simultaneously integrate flux, metabolite, and mRNA data into a functional biological model.
Specifically, our investigation focused on native transcriptional
control of amino acid biosynthetic enzymes via the Gcn4pmediated stress response. This collection of pathways has been
studied extensively by both classical genetics and genomic approaches (3, 12), and human counterparts of Gcn4p-regulated
pathways have been associated with 209 genetic disorders including Phenylketonuria (PKU). Our study aimed to use this
model system to demonstrate that integrating metabolic flux
phenotypes would yield new and specific insights to a wellstudied set of pathways. Here, we present the results of our
measurements, a mechanistic model of amino acid biosynthesis,
a predictor of metabolic flux changes, and a discussion of what
we learned about the metabolic control structure for this model
system.
Results
To study the response enabled by Gcn4p, we compared wild-type
yeast to a gcn4⌬-knockout strain (i.e., stress response present
versus absent) with both strains cultivated in a chemostat under
constant pH, controlled growth rate, and starvation-induced
stress (Fig. 1A). Chemostat cultivations, which enable growth
under a tightly regulated steady-state environment, eliminate
variability that is inherent in dynamically changing batch cultures. In shake-flask studies, for example, it is difficult, and in
some cases impossible, to control several key cultivation parameters (e.g., growth rate, media composition, dissolved oxygen
concentration, pH, etc.), which often obscures the ‘‘real’’ paAuthor contributions: J.F.M., M.C.J., S.G.V.-B., R.T.W., A.G.H., T.I., J.N., and G.S. designed
research; J.F.M., M.C.J., S.G.V.-B., H.A., and L.T. performed research; R.T.W. contributed
new reagents/analytic tools; J.F.M., M.C.J., and M.R.A. analyzed data; and J.F.M., M.C.J.,
M.R.A., S.G.V.-B., and H.A. wrote the paper.
The authors declare no conflict of interest.
This article is a PNAS Direct Submission.
1J.F.M.,
M.C.J., M.R.A., and S.G.V.-B. contributed equally.
2Present
address: Department of Chemical Engineering, University of Delaware, Newark,
DE 19716.
3Present
address: AgResearch Limited, Grasslands Research Centre, Tennent Drive, Private
bag 11008, Palmerston North, New Zealand.
4To
whom correspondence should be addressed. E-mail: gregstep@mit.edu.
This article contains supporting information online at www.pnas.org/cgi/content/full/
0811091106/DCSupplemental.
PNAS 兩 April 21, 2009 兩 vol. 106 兩 no. 16 兩 6477– 6482
APPLIED BIOLOGICAL
SCIENCES
Edited by John Ross, Stanford University, Stanford, CA, and approved February 12, 2009 (received for review November 3, 2008).
A
Environmental Factor
reference
B
i
i
+3AT
2
1
his
3
half lethal
inhibitor
concentration
5
i
Gcn4p
4
His3p
induction
2
1
...
5
HIS3
Gcn4p
13C-label
biom
...
HIS3
wild type
mRNAs
microarrays
HPLC
3
4
His3p
gcn4Δ
GC-MS
his
tyr1
reaction fluxes
(conversion rates)
ass
metabolites
Legend
1) competitive
inhibition
2) enzymatic
activity
3) kinase-activated
translational
regulation
4) transcriptional
activation
5) transcription
and translation
pphn
tyr
independent
phenotypic
measurements,
poorly correlated
due to interactions
like metabolite-enzyme
feedback inhibition
Fig. 1. Experimental design and measurement strategy. (A) Analyzing the
Gcn4p-dependent stress response in a controlled-growth chemostat environment. Wild-type S288C and an isogenic gcn4⌬ derivative were cultured in a
glucose-limited chemostat diluted at 0.10 h⫺1 (416 min per doubling) in
unsupplemented YNB minimal media regulated to 5 ⫹/⫺ 0.1 pH. Titrated
histidine biosynthesis inhibitor levels (10 and 0.1 mM, respectively, of 3-amino1,2,4-triazole or 3AT) created histidine near starvation (Fig. S1), which causes
Gcn4p translational activation and transcriptional activation of hundreds of
targets in the wild type (⫹Gcn4p). Through this modulation of the inhibitor
(3AT) concentration, both wild-type and gcn4⌬ cultures were grown at the
same specific growth rate (0.10 h⫺1) and achieved similar cell densities and
production rates of ethanol and CO2 (Table S3). (B) Multitiered measurement
strategy for analyzing large-scale network perturbations. We used microarrays, GC-MS, and HPLC to measure mRNAs, fluxes, and metabolites (see text
and SI Text for further detail). These independent measurements give a broad
view of the Gcn4p stress response and help to characterize network effects,
such as metabolite interaction densities.
rameter under investigation. This is particularly true when
exploring the transcriptional response, which is known to change
with specific growth rate (13).
Despite similar growth rate conditions and overall macroscopic growth characteristics, the responses of the 2 strains
differed markedly at the molecular level. Our gene expression
measurements (Dataset S1) confirmed that active Gcn4p in the
wild-type population induces transcription of hundreds of target
genes within amino acid biosynthesis and other pathways (3, 14),
as shown in Fig. 2A. The median enzyme mRNA expression level
increased at least 2-fold for the arginine, asparagine, tryptophan,
phenylalanine, tyrosine, histidine, isoleucine, leucine, and valine
pathways, and at least 1 transcript reached this threshold for the
cysteine, methionine, and lysine pathways.
Despite these drastic changes at the molecular level, the wild
type and gcn4⌬ chemostat cultures maintain a macroscopic
physiological similarity. To explain this, we hypothesized that the
cell compensated at the level of metabolism to maintain similar
macroscopic growth under the divergent transcriptional programs observed in the wild-type and gcn4⌬ cultures. Metabolic
compensation can be assessed by monitoring 2 features: metabolite levels— defined as the relative abundances of nongenetically-encoded substrates, intermediates, or products of
metabolic pathways (15), and reaction fluxes—defined as the
rates of conversion by each reaction in the metabolic network
(16) (Fig. 1B). To measure metabolite levels, we used gas
chromatography coupled to mass spectrometry (GC-MS) to
separate, characterize, and quantify metabolites from whole-cell
extracts. Additionally, we used high performance liquid chromatography (HPLC) to obtain more precise measurements of 17
of the 20 free amino acids. In total, we assigned statistical
confidences to abundance ratios of 54 observed metabolites in
the gcn4⌬ versus wild-type cells (Dataset S2).
Reaction fluxes provide a direct measure of metabolite interconversion rates. In the absence of experimental data, values of
reaction fluxes are often estimated in silico under the constraints of
6478 兩 www.pnas.org兾cgi兾doi兾10.1073兾pnas.0811091106
maximized biomass production, an assumed cellular stoichiometry,
a steady state metabolic network and fixed protein composition
(17). However, these simulations result in an underdetermined
system (more fluxes than measurements), and it is still an open
question whether such methods adequately capture regulatoryinduced large network perturbations. For this reason, we cultivated
cells on 1-13C-glucose and experimentally measured conditionspecific biomass amino acid content with HPLC and isotopic
labeling patterns with GC-MS to overdetermine the reaction system
(more measurements than fluxes) and thus more robustly estimate
fluxes (18) that can be found in Dataset S3. Stable isotopic labeling
is an established technique for flux determination whereby metabolic conversion of 13C-enriched substrates generates specifically
labeled metabolites, i.e., isotope isomers, whose labeling patterns
are direct functions of the flux configuration (16). As such, flux can
be estimated from measurements (by GC-MS) of isotopic enrichment of metabolites and solving the inverse problem. Overall, our
measurements yielded 17 amino acid fluxes into biomass and 250
13C-labeling abundances, thus providing a very overdetermined
system (see below) for accurate and robust flux estimation.
To convert these 13C-labeling abundances to intracellular flux
values and statistically verify the results (18), we developed a new
method based on the yeast metabolic reaction network refined
and expanded from Gombert et al. (19). Compared with the 2
existing methods for flux determination (19, 20), we took greater
advantage of redundant flux information (i.e., more measurements than fluxes) by (i) using the complete labeling distributions instead of lumping abundances into a summed fractional
labeling [relative to the GC-MS method (19)], and (ii) simultaneously fitting the entire measurement set [relative to the NMR
method (20)]. We used a network of 83 reactions (with corresponding carbon transitions) distributed among the 3 assigned
compartments (cytosolic, mitochondrial, extracellular), using
75 metabolites (1 substrate, 51 balanced intracellular, and 23
products). Combining our measurement set with this network
resulted in 105 redundant measurements for our now overdetermined system (SI Appendix, C13-Based Reaction Flux Determination). By adding experimentally obtained amino acid
fluxes into biomass measurements (previously assumed unchanging), we expanded our resolution beyond that of previous
methods that only use networks of central carbon metabolism.
We also note the exceptionally high quality of fit in Fig. S2 that
gives us high confidence in our experimentally measured fluxes.
Because metabolite levels and fluxes are independent measurements (21), together they give a more complete view of the
metabolic regulatory system.
Consistent with the previous observation that mRNA expression insufficiently predicts protein level (22), biosynthetic
mRNA changes demonstrated poor correlation with f lux
changes (r ⫽ 0.02 for log ratios or differences in Fig. 2B, r ⫽ 0.07
for deltas or linear differences in Fig. 4B). Indeed, we found little
correlation for any pairwise combination of mRNA, flux, and
end-product metabolite changes (Fig. 2 B–D). Furthermore,
Gcn4p DNA-binding strength did not always translate to transcriptional activation (Fig. 2 A).
We next proposed that a mechanistic model of amino acid
biosynthesis capturing regulatory knowledge would better correlate the mRNA, intracellular flux, and metabolite level datasets. Accordingly, we created a biosynthetic network consisting
of gene/protein, reaction, and metabolite nodes connected by
condition-specific transcription factor binding interactions (3),
protein–protein binding interactions (23), and enzyme-reaction
and reaction-metabolite interactions. This network contained at
its core the metabolic reaction network used previously for (and
statistically verified by) flux determination. Additionally, we
curated all known metabolite-enzyme interactions (24) as an
explicit avenue of enzyme-level regulation. The key feature of
our updated model is that we no longer rely on stoichiometry
Moxley et al.
ecm40
leu9
his5
his3aro3
mRNA log ratio
arg5,6
arg8 arg2
tyr1
lys9
3
metabolite log ratio
pha2
arg3aro8
lys1
ecm40
gdh3
yjl200c
aco1
idh1
gdh1lys12 lys9 lys20 lys2 lys1
gdh3
cit2 idh2
cit3
lys21
lys4 asn1
ilv5gln1
tyr1asn2
leu2 bat2
bat2aro7
ilv3
0.00 cit1
aro9
bat1
ilv3
ilv6
ilv1
ilv2
pro1
pro2
pro3
met14
met3
met17
ecm17
met16
met2
met6
met10
aro8
sul1
sul2
aro1
aro4
aro2
leu1
leu4
bat1
pha2
ilv5
ilv6
ilv2
bat2
bat1
his6
his7
his2 his4
0
-1
-2
-3
r = 0.208
leu9
alt1
asp
r = 0.036
0.25
ecm17
met17
met3
met2
sul2 sul1
2
3
lys
glu
0.00
1
mRNA log ratio
arg
cys
ala
met
gly1
met16
shm2 shm1
met14 met6met10
3
2
thr
0.50
ser33ser2 ser3
gdh1
aat2
pro3sul2
his3 str3
met14
met10 ser1 his4
met16 his7
aat1
thr4
thr1
met17
met2
met3
cys3
str2his5
ecm17
cys4
his1
his2
his6 sul1
pro2
pro1
0
gly
his3
his5
ser2 ser1
0shm2
ser33alt1ser31
shm1
mRNA log ratio
0.75
aro3 leu9
orn
aro8
bat2
bat1
ilv2 ilv3
bat2
bat1
ilv6
ilv1
ilv5bat1 leu4
leu2 bat2 leu1
str3
r = 0.013
alt1
cpa2
2
1
D
cpa2
str2
0.25
-0.25
arg1arg4
ort1
met16
his1
flux log ratio
cpa1
arg1arg4ecm40
hom3
ort1
arg2
met10
cys4
met17
met3
cys3
ecm17
aat2 aat1
sul2
Gcn4p promoter binding log ratio
4
hom2
thr1 arg3
arg8
arg5,6
0.50 met14 met2 sul1
ort1
trp4
arg2
trp3
hom3
trp2 his7
ilv3
gly1 asn1
aro2
aro8aro4
his1
aro1
lys4
his4
met16
ilv2
yjl200c
ilv6
lys1lys2
gln3
asn2
arg8
arg5,6
ser1
1
arg3
trp5
thr1
ilv1
pro2
cit3
ser3gln1
leu4
aat1
lys20 put3
aat2
hom2 arg80
his2 thr4
lys21pha2
rtg3
pyc2
ilv5
tyr1 bat1
leu1
met4
lys9
shm1
ecm17
lys14
alt1ser33 met28
ser2
leu3
sul1
sul2
met10
cys4
cpa1
aro7
met2
idh2
idh1
cys3 pde2
aco1
pde1
mcm1
bat2 met6
0 aro80
his6
met17
cbf1
arg82
leu2
hom6
arg81
pro1 met3
shm2
pro3
met14
met31
lys12
cit1
cit2
pyc1
gcr2
gdh1
met32
gdh3
5
0
1
2
3
4
C
gly1
thr4
hom6
cpa1
r = 0.149
str2
2
0.75
cpa2
arg4
arg1
str3
SPECIAL FEATURE
B
gcn4
aro9
-0.25
-3
gly
-2
ala-1
pro
leu ile
his
val
ser
phe
0
1
2
metabolite log ratio
tyr
3
4
Fig. 2. Pairwise comparisons of changes often reveal poor correlation in Gcn4p binding, mRNA, flux, and end product metabolite. Induction ratios for the
measurements indicated on the axes correspond to Fig. 1. (A) Ratios of mRNAs (measurements described in text) and Gcn4p binding at the upstream promoter under
starvation measured with ChIP arrays as described by Harbison, et al. (3) are scattered. As purely Gcn4p-regulated, CPA2 (Upper Right) illustrates a general observation:
Stronger binding often leads to increased transcriptional activation. For other genes, the binding relationship gives insight to its transcriptional control and identifies
genes (i) bound yet not activated (⬎2 log2 bound, ⬍0.5 log2 expressed) and (ii) activated yet not bound (⬍1 log2 bound, ⬎1 log2 expressed). The first class includes genes
encoding transcription factors Met4p, Leu3p, and Lys14p, which likely require coactivators, and CPA1, which likely undergoes a uORF-mediated activation of mRNA
decay (a metabolite-RNA interaction) (32, 33) by 10-fold increase in free arginine. The second class of genes includes ARO9 that likely have indirect mechanisms of
transcriptional activation, for instance, by Aro80p in response to the 2- and 3-fold increases in free phenylalanine (34). (B–D) Log ratios for the mRNA, reaction flux,
and metabolite measurements described in the text are associated by pathway and labeled. Arginine and aromatic data, which differ in metabolite interaction densities,
are highlighted in red and blue, respectively. Yellow dotted lines highlight the positive x axis, y axis, and diagonal, and the yellow solid line represents the least squares
fit of the data. Data points outside the axes were placed at the periphery. (B) Despite similar transcriptional activations, flux measurements diverge for arginine (in red,
increased) and aromatic (in blue, invariant) pathways. Generally, transcriptional activation is insufficient in predicting flux activation. (C and D) Free phenylalanine and
tyrosine rise 2- and 3- fold upon transcriptional activation, yet aromatic fluxes remains invariant likely because increased product levels modulate pathway activity
through a high density of feedback metabolite-enzyme interactions. Overall, a variety of posttranscriptional control (e.g., metabolite interaction density) mechanisms
simultaneously modulating pathways confound an effort to correlate mRNA, flux, and metabolite changes.
alone. Now, regulatory information (given by interactions within
the network) is incorporated that might impact the correlation
between mRNA levels and metabolic flux. This expanded network model linking all measurement types was visualized in
Cytoscape (25) as shown in Fig. 3 (with a log2 color bar for
measurement ratios of the wild type strain relative to gcn4⌬).
To improve the poor mRNA-flux agreement seen in Fig. 2B, we
used this expanded network model to construct a predictor of
metabolic flux changes as presented in SI Appendix, section 3. A
novelty of the model is the introduction of a parameter termed
‘‘metabolite interaction density’’ for each reaction that captures the
degree to which the reaction’s enzymes are negatively or positively
regulated by metabolites as represented in the above pathway
model. Specifically, for a pathway this quantity is defined as the ratio
of the number of metabolite-enzyme interactions to that of the total
reaction enzymes in the pathway, as displayed in Fig. 4A. Thus,
pathways with high degree of feedback inhibition (Fig. 3 A, B, and
E) and other enzyme-level regulation will have high metabolite
interaction density for their constituent reactions.
Using this concept, network flux changes were predicted from
changes in mRNA levels and metabolite interaction densities,
using the following equation: ⌬flux ⫽ exp(⫺p1 䡠 dinteraction)
Moxley et al.
⌬mRNA
. The rationale for this equation is to allow metabolite
p2
interaction density (dinteraction) to modulate the extent to which
changes in mRNA levels are allowed to propagate to changes in
flux, with p1 and p2 determined to minimize the difference
between actual and predicted fluxes (SI Appendix, mRNA-Flux
Model). According to this equation, the effect of mRNA changes
(⌬mRNA) is attenuated by metabolite interaction density in
causing changes in the pathway flux (⌬flux). The last feature of
our model is the additional constraint that fluxes predicted by
the above equation must also be consistent with the overall
metabolic network stoichiometry and satisfy the steady state
reaction network, i.e., the equation S 䡠 ⌬v ⫽ 0.
This model represents a new, hybrid approach to correlating
mRNA and flux data that combines information from both
interaction network models and flux determination models. A
core element is that the model preserves metabolic network
consistency both in flux determination and in flux prediction
from mRNA data. In addition, the model introduces the new
concept of metabolite interaction density that embodies a hypothesis that a simpler, less-connected regulatory network often
has more correlated mRNA and flux changes.
PNAS 兩 April 21, 2009 兩 vol. 106 兩 no. 16 兩 6479
APPLIED BIOLOGICAL
SCIENCES
3
flux log ratio
A
ser33
ser1
ser3
cys4
ser2
cys3
str2
transcription
factor binding
Node and Edge
Type Key
str3
shm1 shm2
3pg
tf
ser
gly
z1
en
met31
cys
cbf1
mcm1
gly1
A
DN
met32
pep
2
z
en
met28
e
protein-protein
interaction
enzymatic
relationship
1
nz
en
pr
A
z2
e
ot
B
n
tio
ac
re lux
f
in
m
et
o
ab
metabolitee reaction network
metaboliteprotein interaction
lit
met4
sul1
alt1
ala
met17
ecm17 met2
met16 met10
met14
pyr
met3
met6
akiv
ilv3
accoa
Node Color
Key
ilv5
ilv6
ilv1
ilv2
log2 (ratio)
>0.5
>2
0.25
1
reaction
transcript,
0
0
flux
metabolite
-0.25
-1
< -0.5
sul2
< -2
val
ile
pyrm
pde1
met
bat2
hom2
pde2
pyc2
r5p
hom6
pde3
thr
thr4
hser
aat2
asp
aat1
cit3
leu4
B) Ile-Leu-Val,
his6
d=7/11
his3
leu3
his5
his2
cit2
gln3
his
gcr2
rtg3
gcn4
aro1
asn
aro4
aco1
cit
D) Lys,
yjl200c
idh1
akg
his7
cbf1
cit1
fum
his4
leu1
d=6/8
asn1
accoa
oaa
thr1
E) Asp-Thr-Gly,
asn2
leu2
bat1
hom3
pyc1
his1
leu
leu9
idh2
lys20
lys4
gdh1
lys12
aasa
arg2
put3
arg3
arg8
pro3
arg1
trp4
aro8
arg4
ecm40
rtg3
gln3
C) Arg,
aro9
aro80
cpa1
cpa2
trp1
trp5
arg81 mcm1
arg
orn
gln
trp3
aro7
pha2
arg5,6
gln1
chor
trp2
lys
tyr1
pro1
pro
pphn
tyr
glu
pro2
pep
lys1
lys2
hcit
gdh3
aro2
aro3
d=2/7 lys9
lys21
e4p
lys14
ort1
phe
trp
A) Aro,
arg80 arg82
d=1/10
d=13/14
Fig. 3. Integrated perspective of a large-scale network perturbation on amino acid biosynthesis. The central GCN4 node anchors a biomolecular gene/protein
network expanded to include reactions and metabolites. The network contains condition-specific transcription factor binding, protein–protein binding,
enzyme-reaction, reaction-metabolite, and metabolite-protein feedback edges (see node type and edge key). High metabolite interaction density is evident in
A, B, and E. The dotted yellow edges emanating from the GCN4 node indicate an interaction found only in the wild type condition (i.e., not in gcn4⌬)
Measurement ratios of wild type relative to gcn4⌬ were visualized with a log2 colorbar (see node color key), and gray coloring indicates the lack of a measurement
for that node. Findings for individual pathways are discussed in the text.
arg3
orn
glu
arg1
arg
cpa1
arg4
cpa2
ort1
Aromatics
aro4
aro2
aro3
tyr1
aro1
aro8
tyr
aro9
pphn
e4p
phe
chor
pep
aro7
trp
pha2
trp5
trp2
trp1
trp3
trp4
0.75
0.50
0.25
0.00
increasing incorporation of
large-scale network information
predicted delta flux
(from delta mRNA and network)
ecm40
C
m
in et
te ab
r o
de act lite
ns ion
ity
arg8
arg5,6
arg2
scriptional control and gives metabolites an increasing role in
regulating the flux phenotype. This model supports the body of
literature on feedback inhibition in that increased metabolite
1.00
b
ba iom
la a
nc ss
in
g
Arginine
B
m
e
ne tab
tw ol
or ic
k
Metabolite
interaction densities
bi
om m
as RN
s A
bo da
un ta
ds ,
A
transcript-flux agreement
(correlation coefficient)
The model illustrates that flux control (i.e., invariance to perturbation) increases as metabolite—enzyme interaction density
increases. In other words, greater interaction density lessens tran-
12
thr2gly
10
8
6
4
high metabolite
interaction density
chor: 4/4, pphn: 1/1
tyr: 3/4, phe:3/4
lys
hcit aasa
arg
hisval
cys orn
akiv
glu
met ile
leu pphn
chor
tyr
phe pro
2
0
-2
low metabolite
interaction density
orn: 1/3, arg: 0/6
ser
thr
hser
asp
r = 0.801
-4 ser2gly
-6
-4
-2
0
2
4
measured delta flux
6
8
Fig. 4. Incorporating large-scale biomolecular networks improve flux phenotype understanding. (A) Metabolite interaction densities are illustrated for
arginine and aromatics biosythesis. (B) Changes in mRNA measurements were used to predict flux changes as described in the text and SI Text. By adding the
statistically verified reaction network, biomass efflux bounds, and metabolite interaction density, the predicted fluxes from mRNA better matched measured
fluxes with a correlation coefficient value of 0.80 relative to 0.07 corresponding to the initial assumption set (see SI Appendix, section 3). (C) Measured delta flux
for individual reactions is labeled by end product metabolite and plotted versus the flux predicted from mRNA measurements and the 2 parameter model.
Notably, adding metabolite interaction densities discriminates arginine and aromatic pathways to predict arginine’s up-regulated flux. Without the model
incorporating metabolite interaction densities, aromatic pathways would have been wrongly predicted to increase in flux.
6480 兩 www.pnas.org兾cgi兾doi兾10.1073兾pnas.0811091106
Moxley et al.
B
Che mos ta t
e xpe rime nts
75%
Inhibitor
(3AT)
50%
Arg
e nzyme s
S ha ke fla s k
e xpe rime nts
C
∆fluxpre dicte d
Me t4p a nd Cbf1p
pe rturba tions
Thr -> Gly
from ∆mRNA, ne twork
S e r -> Gly
from ∆mRNA, ne twork
Thr -> Gly
from ∆mRNA, ne twork
S e r -> Gly
from ∆mRNA, ne twork
∆me t28
tra ns criptiona l
re gula tion
25%
∆cbf1
0%
wt
∆cbf1
∆me t28
∆flux me a s ure d
a priori mode l pre diction
C13-la be ling flux me a s ure me nts
f (∆mRNA, inte ra ction ne twork) 3 wt x 2 mt e xpe rime nta l compa ris ons
glycine flux
re wire d to a
thre onines ource
Re wiring
pre dicte d?
SPECIAL FEATURE
% Threonine-Originating Flux
A 100%
regulation of enzyme activity results in greater metabolic control.
The high interaction density aromatics and isoleucine-leucinevaline pathways (dinteraction ⫽ 13/14 and dinteraction ⫽ 7/11 in Fig. 3
A and B, respectively) provide an illustrative foil to the low
interaction density arginine and lysine pathways (dinteraction ⫽ 1/10
and dinteraction ⫽ 2/7 in Fig. 3 C and D, respectively) as depicted in
Fig. 4A and discussed in SI Discussion. For the high metabolite
interaction densities, we hypothesize that because end products are
not being increasingly used in protein synthesis, the metabolite
levels build up and cause feedback inhibition at the enzyme level,
resulting in tightly regulated flux. In contrast, we hypothesize that
low metabolite interaction density pathways use alternative strategies. These pathways rely on preferential vacuolar localization of
products and activation of transcription factor regulators separate
from Gcn4p to compensate for unnecessary metabolite build-up.
Despite high metabolite interaction densities of the aspartate and
threonine pathways, dinteraction ⫽ 6/8, fluxes are transcriptionally regulated and increase significantly because of rewiring of glycine flux, as
revealed by the 13C-labeling patterns (Fig. 3E). Glycine may be produced from 3 independently 13C-enriched precursors: glyoxylate
(Agx1p), serine (Shm1p, Shm2p), and threonine (Gly1p). After discarding glyoxylate flux as 0 (within 3% noise) in all conditions, ⬇99%
of glycine flux comes via serine in the reference condition; however, the
presence of activated Gcn4p shifts 44% of that flux to a threonine
precursor. We hypothesize that although free aspartate and threonine
do not increase appreciably, the significant (FDR ⬍2%) activation of
aat2, hom3, and hom2 and near significant (FDR ⬍3% or log2 ratio
⬎1.0) activation of thr1, thr4, and gly1 likely provide an increased
biosynthetic capacity. Thus, some glycine biosynthetic flux is rewired
away from serine to a route depleting the now more available free
aspartate and threonine and increasing fluxes (⬎30% change), i.e., the
up-regulated green reaction nodes in Fig. 3E. Overall, these data
suggest that global rewiring of flux through the network may be a
mechanism to override high metabolite interaction density.
In total, the predictive network model highlights 5 general
observations and 10 new system hypotheses not evident without
high resolution metabolic flux data (Tables S1 and S2). Because
transcriptional flux rewiring has not been studied extensively
and, more specifically, glycine flux rewiring has not previously
been reported as a potential conserved physiological for stress
response, we chose to examine these observations and hypotheses in greater detail.
To validate the predictive network model for transcriptional
changes that drive a rewiring of global network fluxes, we created
Moxley et al.
additional perturbations of yeast transcriptional regulation in the
vicinity of glycine biosynthesis (Fig. 3E) with met28⌬, cbf1⌬,
met31⌬, and met32⌬ knockout strains. We measured mRNA levels
for each strain in shake flask cultures and found that met28⌬ and
cbf1⌬ displayed significant changes in amino acid biosynthetic
pathways (Figs. S3–S6). Using the measured mRNA levels and the
predictive network model from above without any further modification, we determined that both met28⌬ and cbf1⌬ should exhibit
a glycine flux rewiring that favors a threonine precursor.
To experimentally validate the conclusions of the predictive
network model, we used the 13C labeling analysis described above
to measure glycine biosynthetic fluxes for the wild type, met28⌬ and
cbf1⌬ strains in YNB minimal media supplemented with methionine to correct for the corresponding knockout auxotrophies. As
before, we achieved an exceptionally high quality of fit in Fig. S7.
Notably, we observed that the predictive network model correctly identified this experimentally-determined glycine flux rewiring, using the mRNA data. We observed a shifting of glycine flux
to a threonine precursor where serine-originating glycine flux
decreased from ⬇50% in the wild type to ⬇25% in met28⌬ to 0%
in cbf1⌬ as shown in Fig. 5. We hypothesize that the increasing
shutdown of mRNA levels in the methionine pathways increased
the availability of homoserine and thus threonine to yield greater
conversion rates of threonine to glycine.
We were especially encouraged by 3 aspects of the glycine flux
rewiring observations in our follow-on analysis. First, the 13C high
resolution flux determination enabled the experimental observation of a previously unreported mechanism for which an interaction
network allows mRNA to mediate a flux rewiring for a cellular
response. Second, this glycine rewiring mechanism was experimentally observed for perturbations originating in different network
locations indicating a convergent rewiring pathway activated by
transcriptional regulation (Fig. 5B). Finally, we demonstrated that
mRNA and network data may be used to correctly predict flux
changes.
Discussion
Looking forward toward a full fusion of transcriptional and metabolic data, the suspected role of metabolite interaction density and
metabolite-triggered activation of ARO9 transcription (among others) demonstrate opportunities to substantially augment the currently available interactome. Despite modulating nearly all pathways observed, cell-wide data on interactions between metabolites
and proteins are not yet available unlike protein–protein interacPNAS 兩 April 21, 2009 兩 vol. 106 兩 no. 16 兩 6481
APPLIED BIOLOGICAL
SCIENCES
Fig. 5. Observations in glycine flux rewiring demonstrate a biological principle and validate the predictive network model. (A) Additional transcriptional
regulatory perturbations drive a rewiring in glycine flux. Knockout strains for transcriptional regulators in the vicinity of glycine biosynthesis (met28⌬, cbf1⌬,
met31⌬, and met32⌬) were constructed, mRNA levels were measured, and biosynthetic fluxes were determined for the 2 strains (met28⌬ and cbf1⌬) that
displayed significant transcriptional changes. The results for the percentage of threonine-originating glycine flux are displayed in the first panel. Similar to the
previous Gcn4p-induced stress response data, we observe a rewiring of glycine flux away from serine precursor in both the met28⌬ and cbf1⌬ cultures. (B)
Convergent rewiring of glycine flux. In both chemostat and shake flask experiments, transcriptional regulatory networks drives a rewiring of glycine flux to a
threonine source. (C) The predictive model from Gcn4 experiments uses mRNA changes and the interaction network correctly identified flux rewiring in knockout
experiments. mRNA changes from the met28⌬ and cbf1⌬ knockout strains were input to the existing predictive flux model to assess glycine flux rewiring. For
both knockout mRNA profiles, the glycine flux was predicted to be rewired from a serine precursor to threonine. The experimentally measured flux changes of
the serine and glycine (displayed in A) verify the prediction of this observed rewiring.
tions. Chemical genomics investigations screen small molecule
libraries for desired function (e.g., drug efficacy), but the inverse
process to enumerate all interactions of a given metabolite is not as
common. We anticipate that this metabolite-protein interactome
will result from new protein array techniques and iterative interaction network reconstructions of systematically measured fluxes
and metabolite levels akin to previous metabolic and genetic
network reconstructions (26, 27). Just as Hall et al. (28) demonstrated that metabolic enzyme Arg-5,6p was a transcription factor
by globally screening proteins for DNA binding activity, many
metabolite-enzyme interactions remain to be discovered in broad
screens looking outside a targeted system.
In this investigation, we offer evidence that should place bounds
on the interpretation of transcriptional data and provide 5 general
observations and 10 new system hypotheses not evident without
high resolution metabolic data (Tables S1 and S2). We observe that
flux control by metabolites at dense regions of metabolite-enzyme
interactions is a key posttranscriptional regulatory strategy that
allows cells to balance biosynthetic needs for growth. Additionally,
we observe that predictive flux models may be used to correctly
identify flux changes, using mRNA levels and interaction networks.
We conclude that the satisfactory fusion of metabolomic and
fluxomic data for microbial and mammalian systems will not only
benefit from improving analytical and computational techniques
for metabolite and flux measurement but will also necessarily rely
on large-scale interaction networks expanded to include globallyscreened data on metabolite interactions.
L䡠min⫺1, and dissolved oxygen ⬎80% air saturation. The bioreactors were fitted
with cooled condensers (4 °C), and the off-gas was directed to a gas analyzer
(INNOVA) for measurement of CO2 and O2. Steady-state samples were taken after
5 volume changes. Biomass concentration, extracellular metabolite concentrations, and carbon dioxide evolution rate were constant over at least 3 measurements taken before sampling.
Materials and Methods
Chemostat Growth Conditions. Strains were cultivated in aerobic carbon-limited
chemostat cultures in 2-L fermentors (Applikon) in standard YNB media obtained
from Qbiogene with indicated levels of 3-amino-triazole (3AT). The working
volume was 1.0 L and the dilution rate was 0.10 h⫺1. Cultivations were carried out
at 30 °C with an agitation speed of 600 rpm, pH of 5.0, airflow rate of 1.0 standard
ACKNOWLEDGMENTS. We thank Ryan Kelley, Tong Ihn Lee, Scott McCuine,
Chris Workman, Jose Aleman, Richard Young, and Gerald Fink for advice on
experimental design and analysis. This work was supported by National
Institutes of Health Grant 1R01 DK075850-01, the National Science Foundation International Research Fellowship Program, the Singapore–Massachusetts Institute of Technology Alliance, and National Center for Research
Resources award 018627.
1. Uetz P, et al. (2000) A comprehensive analysis of protein–protein interactions in
Saccharomyces cerevisiae. Nature 403:623– 627.
2. Lee TI, et al. (2002) Transcriptional regulatory networks in Saccharomyces cerevisiae.
Science 298:799 – 804.
3. Harbison CT, et al. (2004) Transcriptional regulatory code of a eukaryotic genome.
Nature 431:99 –104.
4. Ideker T, et al. (2001) Integrated genomic and proteomic analyses of a systematically
perturbed metabolic network. Science 292:929 –934.
5. Ideker T, Ozier O, Schwikowski B, Siegel AF (2002) Discovering regulatory and signalling circuits in molecular interaction networks. Bioinformatics 18 (Suppl 1):S233–240.
6. Helmstaedt K, Strittmatter A, Lipscomb WN, Braus GH (2005) Evolution of 3-deoxy-Darabino-heptulosonate-7-phosphate synthase-encoding genes in the yeast Saccharomyces cerevisiae. Proc Natl Acad Sci USA 102:9784 –9789.
7. Hinnebusch AG (1997) Translational regulation of yeast GCN4. A window on factors
that control initiator-trna binding to the ribosome. J Biol Chem 272:21661–21664.
8. Stelling J, Klamt S, Bettenbrock K, Schuster S, Gilles ED (2002) Metabolic network
structure determines key aspects of functionality and regulation. Nature 420:190 –193.
9. Covert MW, Knight EM, Reed JL, Herrgard MJ, Palsson BO (2004) Integrating highthroughput and computational data elucidates bacterial networks. Nature 429:92–96.
10. Fong SS, Nanchen A, Palsson BO, Sauer U (2006) Latent pathway activation and
increased pathway capacity enable Escherichia coli adaptation to loss of key metabolic
enzymes. J Biol Chem 281:8024 – 8033.
11. Kresnowati MTAP, et al. (2006) When transcriptome meets metabolome: Fast cellular
responses of yeast to sudden relief of glucose limitation. Mol Syst Biol 2:49.
12. Hinnebusch AG (1988) Mechanisms of gene regulation in the general control of amino
acid biosynthesis in Saccharomyces cerevisiae. Microbiol Rev 52:248 –273.
13. Fazio A, et al. (2008) Transcription factor control of growth rate dependent genes in
Saccharomyces cerevisiae: A three factor design. BMC Genomics 9:341.
14. Natarajan K, et al. (2001) Transcriptional profiling shows that Gcn4p is a master
regulator of gene expression during amino acid starvation in yeast. Mol Cell Biol
21:4347– 4368.
15. Jewett MC, Hofmann G, Nielsen J (2006) Fungal metabolite analysis in genomics and
phenomics. Curr Opin Biotechnol 191–197.
16. Sauer U (2004) High-throughput phenomics: Experimental methods for mapping
fluxomes. Curr Opin Biotechnol 15:58 – 63.
17. Schilling CH, Palsson BO (1998) The underlying pathway structure of biochemical
reaction networks. Proc Natl Acad Sci USA 95:4193– 4198.
18. Antoniewicz MR, Kelleher JK, Stephanopoulos G (2006) Determination of confidence
intervals of metabolic fluxes estimated from stable isotope measurements. Metab Eng
324 –337.
19. Gombert AK, Moreira dos Santos M, Christensen B, Nielsen J (2001) Network identification and flux quantification in the central metabolism of Saccharomyces cerevisiae
under different conditions of glucose repression. J Bacteriol 183:1441–1451.
20. Maaheimo H, et al. (2001) Central carbon metabolism of Saccharomyces cerevisiae
explored by biosynthetic fractional (13)C labeling of common amino acids. Eur J Biochem 268:2464 –2479.
21. Rossell S, et al. (2006) Unraveling the complexity of flux regulation: A new method
demonstrated for nutrient starvation in Saccharomyces cerevisiae. Proc Natl Acad Sci
USA 103:2166 –2171.
22. Gygi SP, Rochon Y, Franza BR, Aebersold R (1999) Correlation between protein and
mRNA abundance in yeast. Mol Cell Biol 19:1720 –1730.
23. Alfarano C, et al. (2005) The biomolecular interaction network database and related
tools 2005 update. Nucleic Acids Res 33:D418 –D424.
24. Jones EW, Fink GR (1982) in The Molecular Biology of the Yeast Saccharomyces, eds
Strathern JN, Jones EW, Broach JR (Cold Spring Harbor Laboratory, Cold Spring Harbor,
NY), pp 181–289.
25. Shannon P, et al. (2003) Cytoscape: A software environment for integrated models of
biomolecular interaction networks. Genome Res 13:2498 –2504.
26. Vance W, Arkin A, Ross J (2002) Determination of causal connectivities of species in
reaction networks. Proc Natl Acad Sci USA 99:5816 –5821.
27. Gardner TS, di Bernardo D, Lorenz D, Collins JJ (2003) Inferring genetic networks and
identifying compound mode of action via expression profiling. Science 301:102–105.
28. Hall DA, et al. (2004) Regulation of gene expression by a metabolic enzyme. Science
306:482– 484.
29. Antoniewicz MR, Kelleher JK, Stephanopoulos G (2007) Accurate assessment of amino acid
mass isotopomer distributions for metabolic flux analysis. Anal Chem 79:7554 –7559.
30. Yoo H, Antoniewicz MR, Stephanopoulos G, Kelleher JK (2008) Quantifying reductive
carboxylation flux of glutamine to lipid in a brown adipocyte cell line. J Biol Chem
283:20621–20627.
31. Antoniewicz MR, Kelleher JK, Stephanopoulos G (2007) Elementary metabolite units
(EMU): A novel framework for modeling isotopic distributions. Metab Eng 9:68 – 86.
32. Delbecq P, Calvo O, Filipkowski RK, Pierard A, Messenguy F (2000) Functional analysis
of the leader peptide of the yeast gene CPA1 and heterologous regulation by other
fungal peptides. Curr Genet 38:105–112.
33. Gaba A, Jacobson A, Sachs MS (2005) Ribosome occupancy of the yeast CPA1 upstream
open reading frame termination codon modulates nonsense-mediated mRNA decay.
Mol Cell 20:449 – 460.
34. Iraqui I, Vissers S, Andre B, Urrestarazu A (1999) Transcriptional induction by aromatic
amino acids in Saccharomyces cerevisiae. Mol Cell Biol 19:3360 –3371.
6482 兩 www.pnas.org兾cgi兾doi兾10.1073兾pnas.0811091106
Biomass Amino Acid C13 Enrichment Analysis. C13 enrichment biomass was
achieved by chemostat experiments performed in in-house-built reactors with
a working volume of 200 mL. Cultivations were carried out at 30 °C with an
agitation speed of 600 rpm, a dilution rate of 0.1 h⫺1, a pH of 5.0, and airflow
rate of 1.0 standard L䡠min⫺1. Steady-state samples were taken after 5 volume
changes. 100% of the glucose used was labeled in position 1 (1-13C glucose was
from Omicron Biochemicals). The 13C-labeled biomass was harvested by centrifugation at 4,000 ⫻ g and 0 °C for 5 min. After centrifugation, the supernatant was poured off, and the cell pellet was frozen instantaneously in liquid
nitrogen and stored at ⫺80 °C. Hydrolysis, derivatization, and analysis were
carried out as described by Antoniewicz et al. (29).
Metabolic Flux Determination. Metabolic fluxes were determined using the
Metran software, using the elementary metabolite units (EMU) algorithm (30,
31). Metran accepts as input a user-defined metabolic network model consisting of biochemical reactions and atom transitions, and a set of 13Cmeasurements and external fluxes; it produces as output metabolic fluxes for
the entire network, confidence intervals for all fluxes, and statistical analysis
of the goodness-of-fit (18).
Additional materials and methods on strains, shake flask growth conditions, whole genome expression analysis, biomass amino acid composition
analysis, endometabolome analysis, and exometabolome analysis are in SI
Materials and Methods.
Moxley et al.
Download