Construction of large signaling pathways using an adaptive perturbation approach with...

14
This journal is c The Royal Society of Chemistry 2012 Mol. BioSyst., 2012, 8, 1571–1584 1571 Cite this: Mol. BioSyst., 2012, 8, 1571–1584 Construction of large signaling pathways using an adaptive perturbation approach with phosphoproteomic dataw Ioannis N. Melas,z a Alexander Mitsos,z b Dimitris E. Messinis, a Thomas S. Weiss, c Julio-Saez Rodriguez d and Leonidas G. Alexopoulos* a Received 21st November 2011, Accepted 28th February 2012 DOI: 10.1039/c2mb05482e Construction of large and cell-specific signaling pathways is essential to understand information processing under normal and pathological conditions. On this front, gene-based approaches offer the advantage of large pathway exploration whereas phosphoproteomic approaches offer a more reliable view of pathway activities but are applicable to small pathway sizes. In this paper, we demonstrate an experimentally adaptive approach to construct large signaling pathways from phosphoproteomic data within a 3-day time frame. Our approach—taking advantage of the fast turnaround time of the xMAP technology—is carried out in four steps: (i) screen optimal pathway inducers, (ii) select the responsive ones, (iii) combine them in a combinatorial fashion to construct a phosphoproteomic dataset, and (iv) optimize a reduced generic pathway via an Integer Linear Programming formulation. As a case study, we uncover novel players and their corresponding pathways in primary human hepatocytes by interrogating the signal transduction downstream of 81 receptors of interest and constructing a detailed model for the responsive part of the network comprising 177 species (of which 14 are measured) and 365 interactions. Introduction Recent advancements in high throughput technologies are changing the focus of modern biology from the study of individual genes, proteins, or pathways into studying biological systems as a whole. 1 Technologies, such as DNA or protein microarrays for gene expression or protein concentration, Mass Spectrometry (MS) for proteomics, and Yeast two- hybrid (Y2H) screen for Protein–Protein Interactions (PPIs), have led to the construction of vast datasets, addressing cellular behavior on both genomic and proteomic levels. Compilation of all these heterogeneous data into predictive, functional models of cellular processes cannot be tackled by reductionist approaches. 2–4 Instead, markup languages, data annotation, and advanced mathematical modeling are used to make this integration feasible. 5–14 These models are usually depicted in computable pathway maps and used for gaining a deeper understanding of cell’s machinery as well as for identi- fying pathway alterations that take place in complex diseases and cannot be regarded as mere malfunction of a single constituent, that being a gene or a protein. 1 Both gene- and protein-based approaches endeavor the construction of large signaling pathways. A major advantage of the genomic approaches in the pathway construction is the unbiased exploration of the whole genome at the transcrip- tional level. 15–17 However, inconsistencies between protein abundance and corresponding mRNA levels in mammalian cells 18,19 reveal that gene expression alone might not be able to fully explain the function of signaling pathways 1 and thus, proteomic measurements are frequently used for validating gene expression data. On this front, measurement of protein phosphorylation state is considered one of the most reliable surrogates for studying the activity of signaling pathways. 20 However, in contrast to gene expression data, proteomic and phosphoproteomic technologies are far from capable to cover the whole proteome level. 4 Proteomic technologies applied today can be divided into two main categories, the ones that make no a priori assump- tion about the sample’s protein content (e.g. MS), and affinity- based methods (e.g. protein arrays, xMAP technology, aptamers) that rely on antibodies or aptamers to detect a predetermined set of targets. 21 MS and affinity-based methods serve different needs in studying signal transduction networks. MS supports the measurement of thousands of signals, however, it is limited a Dept of Mechanical Engineering National Technical University of Athens, 15780 Zografou, Greece. E-mail: [email protected] b Dept of Mechanical Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139, USA c Center for Liver Cell Research, Department of Pediatrics and juvenile Medicine, University Medical Center Regensburg, Regensburg, Germany d European Bioinformatics Institute (EMBL-EBI), Wellcome Trust Genome Campus, Cambridge CB10 1SD, UK, and European Molecular Biology Laboratory (EMBL), Genome Biology Unit, Meyerhofstrasse 1, D-69117 Heidelberg, Germany w Electronic supplementary information (ESI) available. See DOI: 10.1039/c2mb05482e z These authors contributed equally to this work. Molecular BioSystems Dynamic Article Links www.rsc.org/molecularbiosystems PAPER Downloaded by Massachusetts Institute of Technology on 19/04/2013 20:47:07. Published on 23 March 2012 on http://pubs.rsc.org | doi:10.1039/C2MB05482E View Article Online / Journal Homepage / Table of Contents for this issue

Transcript of Construction of large signaling pathways using an adaptive perturbation approach with...

This journal is c The Royal Society of Chemistry 2012 Mol. BioSyst., 2012, 8, 1571–1584 1571

Cite this: Mol. BioSyst., 2012, 8, 1571–1584

Construction of large signaling pathways using an adaptive perturbation

approach with phosphoproteomic dataw

Ioannis N. Melas,za Alexander Mitsos,zb Dimitris E. Messinis,aThomas S. Weiss,

c

Julio-Saez Rodriguezdand Leonidas G. Alexopoulos*

a

Received 21st November 2011, Accepted 28th February 2012

DOI: 10.1039/c2mb05482e

Construction of large and cell-specific signaling pathways is essential to understand information

processing under normal and pathological conditions. On this front, gene-based approaches offer

the advantage of large pathway exploration whereas phosphoproteomic approaches offer a more

reliable view of pathway activities but are applicable to small pathway sizes. In this paper, we

demonstrate an experimentally adaptive approach to construct large signaling pathways from

phosphoproteomic data within a 3-day time frame. Our approach—taking advantage of the

fast turnaround time of the xMAP technology—is carried out in four steps: (i) screen optimal

pathway inducers, (ii) select the responsive ones, (iii) combine them in a combinatorial fashion to

construct a phosphoproteomic dataset, and (iv) optimize a reduced generic pathway via an

Integer Linear Programming formulation. As a case study, we uncover novel players and their

corresponding pathways in primary human hepatocytes by interrogating the signal transduction

downstream of 81 receptors of interest and constructing a detailed model for the responsive part

of the network comprising 177 species (of which 14 are measured) and 365 interactions.

Introduction

Recent advancements in high throughput technologies are

changing the focus of modern biology from the study of

individual genes, proteins, or pathways into studying biological

systems as a whole.1 Technologies, such as DNA or protein

microarrays for gene expression or protein concentration,

Mass Spectrometry (MS) for proteomics, and Yeast two-

hybrid (Y2H) screen for Protein–Protein Interactions (PPIs),

have led to the construction of vast datasets, addressing

cellular behavior on both genomic and proteomic levels.

Compilation of all these heterogeneous data into predictive,

functional models of cellular processes cannot be tackled by

reductionist approaches.2–4 Instead, markup languages, data

annotation, and advanced mathematical modeling are used to

make this integration feasible.5–14 These models are usually

depicted in computable pathway maps and used for gaining a

deeper understanding of cell’s machinery as well as for identi-

fying pathway alterations that take place in complex diseases

and cannot be regarded as mere malfunction of a single

constituent, that being a gene or a protein.1

Both gene- and protein-based approaches endeavor the

construction of large signaling pathways. A major advantage

of the genomic approaches in the pathway construction is the

unbiased exploration of the whole genome at the transcrip-

tional level.15–17 However, inconsistencies between protein

abundance and corresponding mRNA levels in mammalian

cells18,19 reveal that gene expression alone might not be able to

fully explain the function of signaling pathways1 and thus,

proteomic measurements are frequently used for validating

gene expression data. On this front, measurement of protein

phosphorylation state is considered one of the most reliable

surrogates for studying the activity of signaling pathways.20

However, in contrast to gene expression data, proteomic and

phosphoproteomic technologies are far from capable to cover

the whole proteome level.4

Proteomic technologies applied today can be divided into

two main categories, the ones that make no a priori assump-

tion about the sample’s protein content (e.g.MS), and affinity-

based methods (e.g. protein arrays, xMAP technology, aptamers)

that rely on antibodies or aptamers to detect a predetermined

set of targets.21 MS and affinity-based methods serve different

needs in studying signal transduction networks. MS supports

the measurement of thousands of signals, however, it is limited

aDept of Mechanical Engineering National Technical University ofAthens, 15780 Zografou, Greece. E-mail: [email protected]

bDept of Mechanical Engineering, Massachusetts Institute ofTechnology, Cambridge, MA 02139, USA

cCenter for Liver Cell Research, Department of Pediatrics andjuvenile Medicine, University Medical Center Regensburg,Regensburg, Germany

dEuropean Bioinformatics Institute (EMBL-EBI), Wellcome TrustGenome Campus, Cambridge CB10 1SD, UK, and EuropeanMolecular Biology Laboratory (EMBL), Genome Biology Unit,Meyerhofstrasse 1, D-69117 Heidelberg, Germany

w Electronic supplementary information (ESI) available. See DOI:10.1039/c2mb05482ez These authors contributed equally to this work.

MolecularBioSystems

Dynamic Article Links

www.rsc.org/molecularbiosystems PAPER

Dow

nloa

ded

by M

assa

chus

etts

Ins

titut

e of

Tec

hnol

ogy

on 1

9/04

/201

3 20

:47:

07.

Publ

ishe

d on

23

Mar

ch 2

012

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C2M

B05

482E

View Article Online / Journal Homepage / Table of Contents for this issue

1572 Mol. BioSyst., 2012, 8, 1571–1584 This journal is c The Royal Society of Chemistry 2012

to screening cells under few experimental conditions because

of the semi-automated procedures in sample handling and the

extensive data processing times.22–24 Complementary to MS

approaches, reverse phase protein arrays and xMAP technol-

ogy are confined by a predetermined number of measured

proteins but are capable of measuring thousands of samples in a

single day.25 This advantage—if coupled to assay automation

and automated data gathering—opens the road for studying

cellular pathways in an adaptive fashion in a short time frame.

From the computational standpoint, analysis of phospho-

proteomic data is of major interest in the systems biology

community. On this front, the Dialogue for Reverse Engineering

Assessments and Methods (DREAM3 and 4 conferences)

have crowd-sourced the analysis of phosphoproteomic data

in an effort to identify advantages and limitations of mathe-

matical models for pathway construction.26 Data analysis is

mostly tackled by two classes of methods: data-driven and

topology-driven.27–30 Data-driven analysis is based on methodo-

logies from the field of statistics such as Principal Component

Analysis (PCA), Partial Least Squares Regression (PLSR),

Multiple linear Regression (MR) and clustering, and is applicable

to large signaling pathways.25,31–33 Although such models are

most of the time predictive, they make no use of a priori

knowledge of signaling topology. On the other hand, topology-

driven methods use canonical pathways as a scaffold for building

descriptive models of the signaling network. Formalisms such as

Ordinary/Partial Differential Equations (ODE’s/PDE’s),27,34

Fuzzy Logic and Boolean Modeling35–45 are used to represent

the mechanics of signal transduction. Within the topology-driven

methods, ODE’s are applicable only for small topologies since they

incorporate a large amount of free parameters (rate constants) for

modeling in great detail how signal propagates from one protein to

another. For medium to large topologies Boolean modeling is

currently the method of choice and is used for emphasizing the

structural properties of the pathway (feedback loops, shortest

paths, node dependencies, etc.).36,46–49

As high throughput phosphoproteomic data were made

available, several methods have emerged based on the idea

of optimizing a generic topology to fit the experimental

results.29,50,51 The initial idea on optimization50 lies on a

Boolean representation of a 7-receptor canonical pathway that

is capable of simulating the signal transduction cascade and

compare it to the experimental data. A Genetic Algorithm

(GA) is used to prune the canonical pathway so it best fits a

proteomic dataset and construct a cell-specific pathway. An

alternative approach is introduced in ref. 51 based on an

Integer Linear Programming (ILP) formulation that results

in drastically faster computational times while also providing

an optimality guarantee. The rules of the Boolean logic are

encoded in the constraints of the ILP formulation and solved by

state of the art commercial solvers. The computational effi-

ciency of the ILP approach makes it applicable to optimization

of large topologies. However, for both computational methods,

the quality of the generated pathways relies on the experimental

procedure that should incorporate a very large number of

treatments to constrain the generic topology. In typical phos-

phoproteomic datasets25,51 the treatments include single stimulus

and combination of up to two stimuli (out of B10) and the

pre-treatment with up to two inhibitors (out of B5). It is

understandable that such an experimental procedure is applicable

to small topologies—no larger than 10 receptors—but for large

topologies the number of treatments grows exponentially and

becomes experimentally impracticable.

In this work, we propose a new method for constructing large

topologies based on optimization of canonical pathways with

phosphoproteomic data. On the experimental front we propose

an adaptive approach that is based on the fast turnaround time

of the xMAP technology. Cells are plated on 96-well plates and a

4-step adaptive procedure is employed that includes: (1) ligand

screening, (2) a bimodal model for stimuli selection, (3) design,

implementation, and acquisition of phosphoproteomic data from

combinatorial treatments with selected stimuli in the same batch

of cells, and (4) an ILP formulation for pathway construction

that guarantees a global optimal solution in a computationally

efficient manner. As a case study, we interrogate the signal

transduction network of primary human hepatocytes down-

stream of 81 receptors, which includes all major players of liver

homeostasis but also unknown and less known ligands.

The 4-step approach was implemented in 3-days time frame.

In day 1 the cells are plated. In day 2 the cells are treated with

81 stimuli, phosphoproteomic data acquired, optimal ligand

candidates selected, and new experimental conditions designed.

In day 3 combinatorial treatments for the responsive ligands

are performed, data acquired, and the ILP optimization is

employed. As a result, a detailed, cell-specific signaling path-

way of primary human hepatocytes, incorporating 15 ligands,

177 species and 365 reactions, is constructed within 3 days. To

our knowledge, our approach is a first attempt to construct

large signaling topologies in an adaptive manner by combining

high throughput proteomic measurements with state of the art

optimization algorithms.

Results

Our method employs a 4-step adaptive approach performed

in a 3-day time frame that includes ligand screening, ligand

selection, design and implementation of combinatorial treat-

ments, and pathway optimization. The 4-step procedure is

presented in Fig. 1 and detailed in the following sections.

Ligand screening

Primary human hepatocytes were isolated and plated in 96-well

plates using standard methods as briefly described in Materials

and Methods (day 1).52 A library of 81 stimuli was put together

that consist of all major cytokines, chemokines, and other known

and unknown (i.e. randomly selected) activators of hepatocyte

physiology. Instead of testing the concentration of each individual

stimulus (i.e., by performing a dose response curve and measuring

the phosphorylation levels of several downstream proteins) we

decided to employ a semi-automated text mining approach that

searches the literature and builds a histogram of treated concen-

trations of a particular stimulus (see Materials and Methods).

The concentration histogram provides the user with the most

commonly used concentrations found in the literature and enables

him to choose a high concentration level.

The stimuli library (see the complete names in Table 1)

consists of prototypical players in the medical literature such

as the EGFR pathway (BTC, EGF, TGFA, NRG1, HBEGF),

Dow

nloa

ded

by M

assa

chus

etts

Ins

titut

e of

Tec

hnol

ogy

on 1

9/04

/201

3 20

:47:

07.

Publ

ishe

d on

23

Mar

ch 2

012

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C2M

B05

482E

View Article Online

This journal is c The Royal Society of Chemistry 2012 Mol. BioSyst., 2012, 8, 1571–1584 1573

HGF pathway (HGF), Insulin pathway (INS, IGF1), Toll

Like Receptor Stimuli (LPS, FLAGELLIN, ODN2006,

IMIQUIMOD, HKSA, SSRNA40, OSM, PAM3CSK4,

POLYIC), the BMP family (BMP2, BMP4, BMP7), the TNF

family (TNF, TNFSF10, TNFSF11, TNFSF12, TNFSF14,

TNFRSF11B, TNFSF13B, CD40LG), several inflammatory

and pleiotropic interleukins (IL1A, IL1B, IL2, IL3, IL4, IL6,

IL7, IL8, IL10, IL12, IL13, IL17A, IL19, IL22), lectins

(LGALS1, LGALS3), interferons (IFNB1, IFNG, IFN1 cluster),

inflammatory and other chemokines (CXCL1, CXCL10,

CXCL11, CXCL12, CCL2, CCL4, CCL5, CCL11), the fibroblast

growth factor family (FGF2, FGF4, FGF23), Follistatin-like 1

(FSTL-1), WNT signaling (WNT), as well as several other

inflammatory or pleiotropic activators: INHBA, ADIPOQ,

DEFB1, FCER2, EPGN, EREG, CSF3, GDF5, GDNF,

CSF2, LEP, CSF1, MIA, PDGFB, RLN2, RLN3, TGFB1,

LTA, VEGF121, FST, NOG. High-throughput bead-based

ELISA experiments using semi-automated xMAP techno-

logy (Luminex, Texas, USA) are performed to measure the

‘‘in vitro average phosphorylation level’’ of 14 key phospho-

proteins at 5 and 25 minutes (IKB, JNK, HSP27, CREB,

ERK, STAT3, IRS1S, AKT, GSK3, MAPK14 (i.e. p38),

RPS6KA1 (i.e. p90RSK), MAP2K1 (i.e. MEK1), RSP6KB1/2/3

(i.e. p70S6K), EGFR). The selection of the above-mentioned

signals is based on previous results (14) and assay availability at

the time of this study. The procedure is described in theMaterials

and Methods section and is based on ref. 25 with some minor

modifications. Data were acquired and analyzed the same day

and plotted as shown in Fig. 2 using the DataRail software.53

Ligand selection

66 out of the 81 inducers did not lead to a significant activation

of a key phosphoprotein (Fig. 2a). A logic-based simulation of

Fig. 1 Experimental and computational workflow: (a) cells are plated in 96 well plates. (b) Cells are treated with 81 stimuli. (c) For each stimulus,

the phosphorylation activity of 14 intracellular signals is measured at 5 and 25 minutes post stimulus. (d) Optimal inducers are selected based on a

Gaussian mixture model. (e) The selected inducers are used in a combinatorial fashion. (f) Second phosphoproteomic experiment is performed and

data acquired. (g) The generic pathway is pre-processed via CellNetOptimizer and its (h) observable/controllable part is conserved for

optimization. (i) The pathway is optimized via an Integer Linear Programming formulation.

Dow

nloa

ded

by M

assa

chus

etts

Ins

titut

e of

Tec

hnol

ogy

on 1

9/04

/201

3 20

:47:

07.

Publ

ishe

d on

23

Mar

ch 2

012

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C2M

B05

482E

View Article Online

1574 Mol. BioSyst., 2012, 8, 1571–1584 This journal is c The Royal Society of Chemistry 2012

the generic pathway (see Fig. S1 (ESIw)) such in ref. 36 showed

that in 56 out of the 66 unresponsive stimuli, hepatocytes

should have responded since at least one of the 14 key phospho-

proteins are downstream of the ligand in the generic topology.

For 10 out of the 66 unresponsive stimuli, we cannot observe

their activity because we miss measurements of all downstream

signals (see the following discussion on observability). In order

to select all active ligands, data acquired from the first step were

normalized and a Gaussian Mixture Distribution (GMD)

model was fitted to discretize data to 0 or 1 (OFF or ON

respectively) as detailed in Materials and Methods and plotted

in Fig. 2b. The algorithm leads to the selection of 15 stimuli

based on their strong or marginal activation of at least one key

phosphoprotein, termed ‘‘active responders’’ (Fig. 2c).

Primary human hepatocytes responded to IL6, FLAGELLIN,

TNF, IFNB1, TGFA, TNFSF14, TNFSF12, IL1A, EGF, IL1B,

NRG1, BTC, CD40LG, DEFB1 and LGALS1. As positive

control observations, TGFA, EGF, NRG1 and BTC being

EGFR ligands activated signals related to pro-growth path-

ways, namely AKT, MAP2K1 (MEK), ERK, RPS6KB1,2,3

and RPS6KA1. IL1A, IL1B and TNF activated signals

lying in inflammatory pathways like IKB and HSP27; IL6

activated STAT3. Moreover, less known players were

identified: CD40LG, a member of the TNF superfamily, had

a significant effect on IKB, HSP27, IRS1S and RPS6KB1,2,3;

FLAGELLIN, a TLR5 ligand activated IKB and HSP27;

LGALS1 (also referred as GALECTIN-1) activated RPS6KA1,

RPS6KB1,2,3 and had a medium size effect on STAT3,

IKB and IRS1S; TNFSF12 and TNFSF14 also belonging to

the TNF superfamily yielded medium size responses from

RPS6KB1,2,3, IKB, IRS1S, HSP27 GSK3 and AKT.

This set of ligands together with IFNB1 which in spite of

having medium effects on very few of the signals (STAT3,

RPS6KB1,2,3) is known to play a major role in liver homeo-

stasis, were included in the follow-up experiment.

Design and execution of the combinatorial experiment

In the second phase of the experimental procedure, the 15 stimuli,

found to have strong effects on the 14 key phosphoproteins

in primary hepatocytes, are introduced in a combinatorial

manner in order to build the training dataset that will constrain

the generic topology. The execution of the combinatorial experi-

ment is taking place on the 3rd day of the procedure with the

same batch of primary human hepatocytes plated in day one

Table 1 Stimuli abbreviations. For cytokines, the HUGO nomenclature has been used

Acronym Full name (HUGO nomenclature) Acronym Full name (HUGO nomenclature)

BTC Betacellulin IL22 Interleukin 22EGF Epidermal growth factor LGALS1 Lectin, galactoside-binding, soluble, 1TGFA Transforming growth factor, a LGALS3 Lectin, galactoside-binding, soluble, 3NRG1 Neuregulin 1 IFNB1 Interferon, b1, fibroblastHBEGF Heparin-binding EGF-like growth factor IFNG Interferon, gHGF Hepatocyte growth factor IFN1cluster Interferon, type 1, clusterINS Insulin CXCL1 Chemokine (C–X–C motif) ligand 1IGF1 Insulin-like growth factor 1 CXCL10 Chemokine (C–X–C motif) ligand 10LPS Interferon regulatory factor 6 CXCL11 Chemokine (C–X–C motif) ligand 11FLAGELLIN Component of the bacterial flagellar filament CXCL12 Chemokine (C–X–C motif) ligand 12ODN2006 Synthetic oligonucleotides CCL2 Chemokine (C–C motif) ligand 2IMIQUIMOD Imidazoquinoline amine analogue to guanosine CCL4 Chemokine (C–C motif) ligand 4HKSA Preparation of Listeria monocytogenes CCL5 Chemokine (C–C motif) ligand 5SSRNA40 20-mer single-stranded RNA oligo CCL11 Chemokine (C–C motif) ligand 11OSM Oncostatin M FGF2 Fibroblast growth factor 2 (basic)PAM3CSK4 Synthetic tripalmitoylated lipopeptide FGF4 Fibroblast growth factor 4POLYIC Synthetic analog of double-stranded RNA FGF23 Fibroblast growth factor 23BMP2 Bone morphogenetic protein 2 FSTL1 Follistatin-like 1BMP4 Bone morphogenetic protein 4 WNT Wingless-type MMTV integration site family, mem. 1BMP7 Bone morphogenetic protein 7 INHBA Inhibin, beta ATNF Tumor necrosis factor ADIPOQ Adiponectin, C1Q and collagen domain containingTNFSF10 Tumor necrosis factor lig. superfamily, #10 DEFB1 Defensin, beta 1TNFSF11 Tumor necrosis factor lig. superfamily, #11 FCER2 Fc fragment of IgE, low affinity II, receptor for CD23TNFSF12 Tumor necrosis factor lig. superfamily, #12 EPGN Epithelial mitogen homologTNFSF14 Tumor necrosis factor lig. superfamily, #14 EREG EpiregulinTNFRSF11B Tumor necrosis factor rec. superfamily, #11b CSF3 Colony stimulating factor 3 (granulocyte)TNFSF13B Tumor necrosis factor lig. superfamily, #13b GDF5 Growth differentiation factor 5CD40LG CD40 ligand GDNF Glial cell derived neurotrophic factorIL1A Interleukin 1 a CSF2 Colony stimulating factor 2IL1B Interleukin 1 b LEP LeptinIL2 Interleukin 2 CSF1 Colony stimulating factor 1 (macrophage)IL3 Interleukin 3 MIA Melanoma inhibitory activityIL4 Interleukin 4 PDGFB Platelet-derived growth factor a polypeptideIL6 Interleukin 6 RLN2 Relaxin 2IL7 Interleukin 7 RLN3 Relaxin 3IL8 Interleukin 8 TGFB1 Transforming growth factor, beta 1IL10 Interleukin 10 LTA Lymphotoxin a (TNF superfamily, member 1)IL12 Interleukin 12 VEGF121 Vascular endothelial growth factorIL13 Interleukin 13 FST FollistatinIL17A Interleukin 17 a NOG NogginIL19 Interleukin 19D

ownl

oade

d by

Mas

sach

uset

ts I

nstit

ute

of T

echn

olog

y on

19/

04/2

013

20:4

7:07

. Pu

blis

hed

on 2

3 M

arch

201

2 on

http

://pu

bs.r

sc.o

rg |

doi:1

0.10

39/C

2MB

0548

2E

View Article Online

This journal is c The Royal Society of Chemistry 2012 Mol. BioSyst., 2012, 8, 1571–1584 1575

and stored in the incubator for one additional day to be used

in this 2nd phase of experiments.

As mentioned in previous studies,30 the construction of

optimal topologies entails biases due to (i) experimental design,

(ii) the signals that are selected to be measured, and (iii) the

number of perturbations via inhibitors imposed in the network.

Finding an optimal experimental design formaximally constraining

a generic topology is not a trivial task. Pathway controllability

Fig. 2 Ligand selection: (a) 81 stimuli are screened and data acquired. (b) A Gaussian mixture model is used to classify signals between activated

and inactivated. (c) Ligands that activated at least one of the measured key phosphoprotein signals are selected.

Dow

nloa

ded

by M

assa

chus

etts

Ins

titut

e of

Tec

hnol

ogy

on 1

9/04

/201

3 20

:47:

07.

Publ

ishe

d on

23

Mar

ch 2

012

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C2M

B05

482E

View Article Online

1576 Mol. BioSyst., 2012, 8, 1571–1584 This journal is c The Royal Society of Chemistry 2012

and observability,50 assay cost limitations, assay performance,

reagent performance, and experimental compatibility with

12 � 8 well plate layout should be taken into consideration.

Based on previous assessment of model sensitivity to changes

in experimental design,30 in this study we created a dataset that

is experimentally feasible and includes all single treatments,

the majority of combination of two together, and the presence

or absence of two selective inhibitors, MEKi and PI3Ki (see

the Phospho-proteomic experiments section in Materials and

Methods). Although combinations of more than two ligands

may uncover more complex cross-talks and synergistic effects,

if done rigorously, the number of experiments needed to

include all the different combinations would increase signifi-

cantly, consequently increasing both the experimental and

computational costs. Instead of incorporating more combina-

torial patterns we chose to increase the number of interrogated

ligands, since the main purpose of this paper is the broad,

qualitative, less detailed study of the signaling network in

human hepatocytes.

The complexity of the combinatorial dataset in Fig. 3 offers

little potential for manual inspection. Nevertheless, a few basic

trends can be identified. MEKi and PI3Ki have blocked their

nominal targets: the activation of ERK is clearly inhibited by

MEKi under any treatment and PI3Ki blocks the activation of

AKT under any treatment. The single treatments included in

the second dataset are grouped together and aligned with the

respective treatments from the first dataset, so a direct com-

parison can be made: as positive controls, all the pro-growth

stimuli have activated AKT, MAP2K1 (MEK), ERK,

RPS6KB1,2,3 and RPS6KA1, as in the first dataset and the

pro-inflammatory ligands activated MAPK14, HSP27 and JNK.

Despite the fact that qualitatively most ligands performed the

same in the 1st and 2nd dataset, few quantitative differences

can also be observed: (i) the JNK and MAPK14 signals clearly

respond to IL1A, IL1B and TNF in contrast to the first

dataset that shows marginal—if any—JNK and MAPK14

activation. (ii) TGFA, EGF and BTC show strong activation

in both phases but with some increased activation levels in

some of the signals in the second phase (MAPK14, JNK,

HSP27). From our experience, mismatches in measurements

of this kind are common even when both experiments are

performed under the exact same conditions, with the same

batch of reagents, and on the same batch of cells from the

same donor. What is important for the optimization process is

the encapsulation of the clear trends of the signals, since small

levels of noise are countered by internal replicates of the

dataset and the normalization procedure.

Clustering

Clustering algorithms (see Materials and Methods) can be

utilized to provide powerful insight when such a large panel of

ligands is screened. By evaluating the Euclidean distance (in <14)

amongst the 81 stimuli, we identify the ones that exhibit

similar profiles to prototypical anabolic or catabolic behavior.

The second dataset is clustered hierarchically respective to

both axes (stimuli and signals). As far as the stimuli axes are

concerned, they seem to be separated into 2 main clusters,

the 1st cluster includes IL6, BTC, EGF, TGFA and NRG1,

the 2nd cluster includes CD40LG, TNF, IL1B, IL1A and

FLAGELLIN, while IFNB1, FSTL1, LGALS1, TNFSF12,

TNFSF14 yield marginal response, therefore are left out of all

clusters and the analysis that follows. In a similar fashion

signals are separated into 2 main clusters. The first one consists

of EGFR, ERK, MAP2K1, RPS6KA1, STAT3, IRS1S, AKT,

RPS6KB1/2/3 and GSK3. The second cluster consists of IKB,

MAPK14, HSP27 and CREB. Interestingly, the ligands of the first

cluster share some pro-growth qualities by activating signals in the

first cluster (pro-growth signals), whereas the second stimuli cluster

consisting of pro-inflammatory stimuli activates mostly the second

signal cluster (pro-inflammatory signals) (see also Fig. 3). Such a

procedure reveals the inflammatory role of FLAGGELIN and

CD40LG, two less-known stimuli that cluster closely to proto-

typical inflammatory ones. For a clustering analysis of the full

combinatorial dataset see S2 and Fig. S3 (ESIw).

Construction of generic pathway

The generic pathway presented in Fig. 1 (and in more detail in

Fig. S1, ESIw) is constructed downstream of 81 receptors of

interest and in the neighborhood of 14 measured proteins.

Several online databases were queried, but most of the reactions

were obtained from KEGG (http://www.genome.jp/kegg/) and

Ingenuity (http://www.ingenuity.com/). Conflicting reports on

protein interactions were handled by manual search of

PubMed database and selection of the most cited alternative.

Addressing the problem of combining pathways from various

databases, the HUGO Gene Nomenclature Committee database

(http://www.genenames.org) was used to ensure standardized

protein names. The pathway was built and visualized using

Graphviz (http://www.graphviz.org/) in a similar fashion to

ref. 50 following manual curation. The resulting topology is

made of more than 500 species and 1000+ reactions (see

Fig. 1g and Fig. S1, ESIw).

Pathway pre-processing: controllability, observability and

feedback loops

Out of the 81 stimuli interrogated originally, only 15 were

found to have strong effects on any of the key phosphorylation

signals. Thus, a large portion of the generic pathway can be

removed before the optimization procedure that is either non-

perturbed (by the 15 cytokines) or non-observable.50 Using

Warshall’s algorithm for transitive closure, as implemented in

CellNetOptimizer (CellNOpt) toolbox,50 the connectivity of

every node to the 14 signals and 15 cytokines is examined (see

Material and Methods). If a pathway exists leading from at

least a perturbed receptor to an arbitrary node A, then A is

considered controllable. If a pathway exists leading from A to

at least a signal, then A is considered observable. The observable

and controllable subset of the generic topology is plotted in

Fig. 1 (and in more detail in Fig. S2, ESIw) numbers 177

species and 365 reactions and is the network to be optimized

by the ILP. Note that the reduction of the network also results

in a significant reduction of the optimization formulation

which makes it computationally more tractable.

The generic topology numbers a large amount of feedback

loops that have to be identified before the optimization

procedure. The presence of a feedback loop implies that

Dow

nloa

ded

by M

assa

chus

etts

Ins

titut

e of

Tec

hnol

ogy

on 1

9/04

/201

3 20

:47:

07.

Publ

ishe

d on

23

Mar

ch 2

012

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C2M

B05

482E

View Article Online

This journal is c The Royal Society of Chemistry 2012 Mol. BioSyst., 2012, 8, 1571–1584 1577

species can affect their own activation state. ODEs (or other

frameworks for dynamic modeling) support this kind of

dependencies, since the feedback occurs in a subsequent time

point. The Boolean approach used herein utilizes only ‘‘early

response data’’ and thus assumes only a feed-forward approach

where feedback loops are not considered. Thus, the feedback

loops present in the generic topology are identified using an

in-house Depth-First Search algorithm and additional con-

straints are crafted to prohibit the ILP from including them in

the solution (see S3, ESIw).

Fig. 3 Combinatorial experiment: (a) design of the combinatorial experiment. The illustrated stimuli treatments are repeated 3 times: without any

inhibitor present, with a MEK inhibitor (MEKi), and with a PI3K inhibitor (PI3Ki). (b) Phosphoproteomic data are acquired acquisition and

plotted in DataRail.

Dow

nloa

ded

by M

assa

chus

etts

Ins

titut

e of

Tec

hnol

ogy

on 1

9/04

/201

3 20

:47:

07.

Publ

ishe

d on

23

Mar

ch 2

012

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C2M

B05

482E

View Article Online

1578 Mol. BioSyst., 2012, 8, 1571–1584 This journal is c The Royal Society of Chemistry 2012

Pathway optimization

The optimization procedure is built around an Integer Linear

Programming formulation, that receives as input a generic

topology and a phosphoproteomic dataset, and by altering the

topology that contradicts experimental observations delivers a

signaling pathway that fits the experimental dataset with the

least possible discrepancies.51 Pathway optimization relies on

the minimization of a two term objective function, the first

term penalizes the experiments-topology mismatch and the

second term either penalizes or rewards the size of the path-

way30,50 (see also ILP Formulation in Materials and Methods).

Penalty weights are adjusted by assuming very small values so

that the goodness of fit is always prioritized over the size of the

pathway.30 In contrast to our previous procedures50,51 that

utilize the minimization of pathway size as a method to narrow

down the solution pool, we consider additional pathway

solutions of larger size that may bear strong biological signi-

ficance and should be conserved. Therefore, in the current

analysis we apply the ILP formulation under 2 different

settings: first by using the standard procedure with small

positive size weights, we minimize the size of the pathway;

second by using small negative reaction weights we maximize

the size of the pathway while retaining the prioritization over the

goodness of fit. The dual-settings approach we adopt aims to

obtain a superset of all possible pathway solutions, i.e., pathways

with more nodes and edges but with the same match of the

experimental data. Additionally, we provide an indicative solution

with the minimum number of reactions. The ILP formulation was

solved by GUROBI (Gurobi library version 3.0.1. Houston,

Texas: Gurobi Optimization, Inc., http://www.gurobi.com/)

through GAMS54 and the optimization results are illustrated in

Fig. 4. Dashed lines correspond to the ‘‘negative weights’’-solution

and the bold ones to the minimum-size solution. For a better

assessment of the ILP performance and differences between

positive and negative size weights see also S4 (ESIw).Out of the original 365 reactions present in the reduced

canonical pathway, 204 were removed by the ILP formulation

(Fig. 4) for the maximum-size solution (superset has 161

reactions—Fig. 4, ‘‘dashed lines’’). In the minimum-size

solution only 53 reactions are included. As positive control

observations, the modular activation patterns described pre-

viously are conserved. The pro-growth stimuli, namely EGF,

TGFA, NRG1, BTC and IL6, signal through similar, partially

overlapping pathways and activate the first signal cluster

consisting of EGFR, AKT, MAP2K1, ERK, STAT3 and

RPS6KA1. On the other hand TNF, CD40LG, IL1A, IL1B

and FLAGELLIN signal through TRAF6, TRAF2 and the

MAPKs and activate MAPK14, IKB, JNK, HSP27 and

CREB. RPS6KB1/2/3, GSK3 and IRS1S, although seemed

to respond to pro-growth stimuli, are removed from the

pathway, since their basal level is relatively high implying a

low signal to noise ratio and causing the fold change to drop

below 2 (that is the threshold for considering a signal activated,

see also ref. 17 and 24 and Materials and Methods section). In

more detail, concerning the EGFR pathway: NRG1, BTC,

EGF and TGFA signal viaGRB2 to RAS and then to (i) PI3K-

AKT and to (ii) RAF - MAP2K1 - ERK - RPS6KA -

CREB, ERK - STAT3. Another branch downstream of IL6

signals via JAK1 to (i) PI3K - AKT and to (ii) STAT3.

Concerning the pro-inflammatory ligands (IL1A, IL1B,

FLAGELLIN, CD40LG and TNF): IL1A, IL1B and

FLAGELLIN signal through TRAF6 to MAP3K7 and then

to (i) NIK- IKB, and to (ii) MAP2K4- JNK,MAP2K4-

MAPK14, MAP2K4 - MAPK11 - HSP27, MAPK11 -

RPS6KA4 - CREB. CD40LG signals through TRAF5 to

NIK and activates IKB. TNF signals through TRAF2 to

MAP3K7 and activates MAPK14, CREB, HSP27, JNK and

IKB. It is clear that minimum-size solution (bold lines) and

superset of possible solutions (dashed lines) serve the exact

same functionality in terms of connecting stimuli to signals.

Their sole difference is that the superset of possible solutions

includes, apart from the shortest, all alternative paths for

achieving that connectivity.

The optimization procedure caused the fitness error to drop

from 31% to 7%, validating that the canonical pathway alone

could not capture the signal transduction mechanisms of the

specific cell type. The ILP formulation managed to fit the basic

trends of the experimental dataset and construct a hepatocyte-

specific signaling pathway. See S5 (ESIw) for a more detailed

inspection of fitness error, before and after the optimization

procedure.

Conclusions

In this manuscript, we propose an approach capable of

interrogating signal transduction downstream of 81 receptors

of interest and subsequently, constructing a detailed view of

the signaling pathway for the responsive part of the network,

numbering 15 receptors, 177 species and 365 reactions, all

within a 3-day time frame. In contrast to MS-based techniques,

the fast turnaround time of the xMAP technology allowed an

adaptive approach that learns from a functional phospho-

proteomic screen to define the most appropriate combinatorial

experiments for pathway construction. The 3-day approach is

performed on the same batch of cultured primary cells, thus

reducing cell differentiation and eliminating donor-to-donor varia-

bility. On the computational front, the ILP formulation solved by

Gurobi, a state of the art commercial solver, has successfully

negotiated the optimization of a canonical pathway to best fit the

functional characteristics of the interrogated cell line.

In contrast to most approaches applied today that focus on

the well-studied and the most-cited pathways, in this project

we rely on a functional screen of a large library of stimuli to

select the most appropriate inducers of hepatocyte intracellular

activity. As expected, most ligands we found correspond to

well-known and highly cited ‘‘prototypical’’ players for liver

homeostasis but surprisingly, we also uncover poorly and

unreported stimuli that induced strong intracellular responses.

Clustering analysis combined with literature search was able to

uncover the potential role of those stimuli due to their func-

tional similarity with ‘‘prototypical’’ ones. On the other hand,

our literature-unbiased approach for ligand selection cannot

identify all active players for 4 main reasons: (i) it is based on

key measured signals that might not belong to the downstream

cascade of the activator (e.g., WNT signaling could not be

detected due to the absence of a beta-Catenin signal55); (ii) the

phosphorylation events are below the limits of detection of

Dow

nloa

ded

by M

assa

chus

etts

Ins

titut

e of

Tec

hnol

ogy

on 1

9/04

/201

3 20

:47:

07.

Publ

ishe

d on

23

Mar

ch 2

012

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C2M

B05

482E

View Article Online

This journal is c The Royal Society of Chemistry 2012 Mol. BioSyst., 2012, 8, 1571–1584 1579

the assay; (iii) the activation occurs in a different time point than

the one measured (i.e. not at 5 and 25 minutes). (iv) Finally,

since ligands are screened based on their individual effects on

the measured proteins, synergistic effects cannot be captured

at this point. Implying that players may be left out of the

following steps of the analysis with strong, yet unidentified,

synergistic effects. In fact, components showing only synergistic

effects are lost at this point. As more phosphoprotein assays

Fig. 4 Optimized pathway: the optimized pathway consists of the reactions and nodes conserved by the ILP formulation.

Dow

nloa

ded

by M

assa

chus

etts

Ins

titut

e of

Tec

hnol

ogy

on 1

9/04

/201

3 20

:47:

07.

Publ

ishe

d on

23

Mar

ch 2

012

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C2M

B05

482E

View Article Online

1580 Mol. BioSyst., 2012, 8, 1571–1584 This journal is c The Royal Society of Chemistry 2012

are added in the future and more time points are selected,

more activators can be identified.

Based on our ligand screening approach we identified two

under-reported compounds: Flagellin and CD40LG. Flagellin,

a TLR5 inducer, was found to activate pro-inflammatory

pathways, and function—as the clustering analysis revealed—

in a similar way to TNF, IL1A and IL1B; enforcing the

hypothesis that hepatocytes assume an active role in immuno-

surveillance mechanisms.25,56 Moreover, CD40LG, a member

of the TNF superfamily previously reported to be expressed in

T-cells and affect B-cells and endothelial cells, is identified as

having a clear effect on primary hepatocytes by activating the

NFkB pathway. CD40LG induced activation of IKB and P38/

HSP27 cluster closely resembling the prototypical TNF, IL1B

and IL1A inflammatory players, validating its role in liver

inflammation. Even though the role of CD40L in hepatocyte

physiology is not clear from the literature, a single study by

Zhou et al.57 identifies CD40 as a driving factor in the patho-

genesis of fulminant hepatitis via its upregulation in Kupffer

cells and hepatocytes. Together with the present results,

CD40L should be considered a major inflammatory player

in liver physiology. Apart from the identification of novel

players, the high throughput screening revealed clustering of

the stimuli in two main groups, the first one consisting of

EGF, TGFA, NRG1, BTC and IL6, sharing some pro-growth

qualities and the second one consisting of TNF, IL1A, IL1B,

FLAGELLIN and CD40LG activating mostly pro-inflammatory

pathways (IKB, MAPK14, HSP27, JNK). As expected, this

modularity has led to the construction of partially overlapping

pathways, namely, EGF, TGFA, BTC and NRG1 all signal

through the EGFR and GRB2 adaptor protein, while IL1A,

IL1B and FLAGELLIN signal through the IL1R and IRAK.

The ILP formulation conserved a subset of the reactions

found in the canonical pathway, shaping a descriptive model

of how signal propagates in primary hepatocytes. However,

the construction of Boolean models of signaling pathways

from phosphoproteomic data entails several validation issues

related to optimal experimental design, model sensitivity to the

generic topology, and calibration of the weights of the two

objective terms (detailed description can be found in ref. 30

and 50). The major limitations in pathway optimization rely

on two main factors: pathway controllability (e.g., how many

stimuli and downstream inhibitors can be used), pathway

observability, and experimental design. Amongst all, the number

of the key phosphoprotein signals is of utmost importance for

increased pathway observability. Surprisingly, just 14 phospho-

protein signals used in this study were sufficient to give a

pathway coverage equal to 68.5% of the generic (685 out of

the 1000 reactions can be observed from the generic topology

when all stimuli are used). Further development of high

throughput assays will enable us to increase the pathway

coverage but also increase the quality of results since more

than one signal in the same pathway increase the reliability of

the pathway topology.

Construction of predictive models for the intracellular

signaling cascades is the cornerstone for understanding cellular

behavior. Here we presented an integrative approach to construct

large scale signaling pathways, based on an adaptive collection

of high throughput phosphoproteomic data and a priori

knowledge of protein connectivity. State of the art optimiza-

tion algorithms were coupled to an ILP optimization formula-

tion to prune a canonical pathway of 177 nodes and 365

reactions to best fit the dataset at hand. Our approach sheds

light into the complex signaling mechanisms of primary human

hepatocytes and constitutes a proof-of-principle for construc-

tion of large pathways based on combinatorial data and a

limited number of measured signals.

Methods

Hepatocyte isolation and plating

Liver tissue samples are obtained from patients undergoing

partial hepatectomy for metastatic liver tumor secondary to

colorectal cancer. Tumor aggregates are resected including a

safety margin within the normal tissue and histological exami-

nation by a pathologist confirms that tumor-free liver tissue is

obtained for cell isolation. Procedures are performed using

protocols approved by the local ethical committee and according

to the guidelines of the charitable state-controlled foundation

Human Tissue and Cell Research with the patient’s informed

consent.58 Primary human hepatocytes were isolated and the

cell viability was examined by Trypan blue exclusion.52 If cell

preparations had a viability of more than 80% they were

plated in collagen type I-coated 96 well plates at a density of

1.2 � 105 cells cm�2 and cultivated in Williams Medium E

(Biochrom AG, Berlin, Germany) supplemented with 2 mM

l-glutamine (Invitrogen), 100 nM dexamethasone (Sigma) and

1% penicillinstreptomycin (Invitrogen).

Determining treated concentrations of stimuli

A custom, semi-automated, literature mining approach was

implemented for determining the saturated levels of the 81

stimuli in the library. Google search engine was queried with

the stimuli name (and common aliases found in HUGO Gene

Nomenclature Committee database) preceded by ‘‘mg/ml’’,

‘‘ng/ml’’, ‘‘mM’’, ‘‘mM’’, ‘‘nM’’, etc. and each time the first 100

results were reviewed. The system searches the first 100 results

of Google Scholar for the preferred concentration format of

the search term, results are stored and the numeric results are

filtered. Then the system evaluates the frequency of appear-

ance for the different concentrations.

The stimuli concentrations that hepatocytes were treated

with are: ADIPOQ (100 ng ml�1), TNFSF13B (25 ng ml�1),

DEFB1 (100 ng ml�1), BMP2 (100 ng ml�1), BMP4 (100 ng ml�1),

BMP7 (100 ng ml�1), FCER2 (100 ng ml�1), CD40LG (1000

ng ml�1), CTGF (50 ng ml�1), CCL11 (100 ng ml�1), FGF2

(50 ng ml�1), FGF4 (50 ng ml�1), FGF23 (50 ng ml�1), FST

(100 ng ml�1), CSF3 (50 ng ml�1), GDF5 (100 ng ml�1),

GDNF (100 ng ml�1), CSF2 (50 ng ml�1), IFNA1 (100 ng ml�1),

IFNB1 (100 ng ml�1), IL29 (100 ng ml�1), IL3 (50 ng ml�1),

IL10 (50 ng ml�1), IL12 (50 ng ml�1), IL13 (50 ng ml�1), IL19

(100 ng ml�1), IL22 (100 ng ml�1), CXCL11 (100 ng ml�1),

LEP (100 ng ml�1), CSF1 (50 ng ml�1), MIA (500 ng ml�1),

NOG (200 ng ml�1), TNFSF11 (50 ng ml�1), RLN2

(100 ng ml�1), RLN3 (100 ng ml�1), CXCL12 (100 ng ml�1),

LTA (100 ng ml�1), TNFSF10 (100 ng ml�1), TNFSF12

(100 ng ml�1), VEGFA (100 ng ml�1), IGFBP1 (100 ng ml�1),

Dow

nloa

ded

by M

assa

chus

etts

Ins

titut

e of

Tec

hnol

ogy

on 1

9/04

/201

3 20

:47:

07.

Publ

ishe

d on

23

Mar

ch 2

012

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C2M

B05

482E

View Article Online

This journal is c The Royal Society of Chemistry 2012 Mol. BioSyst., 2012, 8, 1571–1584 1581

IGFBP3 (100 ng ml�1), LGALS1 (2000 ng ml�1), LGALS3

(2000 ng ml�1), TNFRSF11B (100 ng ml�1), WNT1

(100 ng ml�1), Pam3CSK (1000 ng ml�1), HKLM (108 c ml�1),

PolyIC (10 000 ng ml�1), LPS (10 000 ng ml�1), Flagellin

(1000 ng ml�1), FSL1 (1000 ng ml�1), Imiquimod

(1000 ng ml�1), ssRNA (1000 ng ml�1), ODN2006 (5 mM),

TGFA (200 ng ml�1), BTC (50 ng ml�1), EGF (100 ng ml�1),

EPGN (100 ng ml�1), EREG (100 ng ml�1), CXCL1

(100 ng ml�1), HBEGF (50 ng ml�1), NRG1 (100 ng ml�1),

HGF (100 ng ml�1), IFNG (100 ng ml�1), Igf1 (100 ng ml�1),

IL1A (20 ng ml�1), IL1B (20 ng ml�1), IL2 (50 ng ml�1), IL4

(50 ng ml�1), IL6 (100 ng ml�1), IL7 (100 ng ml�1), IL8

(100 ng ml�1), IL17A (100 ng ml�1), CXCL10 (100 ng ml�1),

TNFSF14 (100 ng ml�1), CCL2 (100 ng ml�1), CCL4

(100 ng ml�1), OSM (100 ng ml�1), PDGF-BB (100 ng ml�1),

CCL5 (100 ng ml�1), TGFB1 (10 ng ml�1), TNF (100 ng ml�1),

INS (1720 ng ml�1).

Phospho-proteomic experiments

Phosphoproteomic measurements were carried out using

bead-based ELISA type assays (xMAP technology). The

following phosphoprotein bead set from Bio-Rad were used:

IkB-a (Ser32/Ser36), JNK (Thr183/Tyr185), HSP27 (Ser78),

CREB (Ser133), ERK1/2 (Thr202/Tyr204, Thr185/Tyr187),

STAT3 (Ser727), IRS1Ser (Ser636/Ser639), Akt (Ser473),

GSK3(Ser21/Ser9), MAPK14 (i.e. p38) (Thr180/Tyr182),

RPS6KA1 (i.e. p90RSK) (Thr359/Ser363), MAP2K1

(i.e. MEK1) (Ser217/Ser221), RSP6KB1/2/3 (i.e. p70S6K)

(Thr421/Ser424). The EGFR signal (total tyrosine) was

measured using the Millipore assay. The signals are chosen

based on the availability of the reagents and quality controls

performed at the early phases of the experimental setup.

Originally, 20 signals were measured, including (apart from

the 14 mentioned above) BCR245 (B-Cell Receptor), IRB

(Insulin Receptor B), PDGFR, SRC, CMET and IRS1T.

However, high basal levels and experimental noise led to our

decision to exclude them from the dataset. The experimental

protocol25 was altered and the incubation times were set to

1 hour for the samples/beads treatment, 2 hours for the

secondary antibody, and 10 minutes for the PE in order to

minimize the total experimental time and make it compatible

with the adaptive procedure.

At the second day of the procedure (after the hepatocyte

plating in day 1) each of the 81 stimuli was applied in each well

and cells were treated in separate plates for 5 and 25 minutes.

The selection of these two time points was based on preliminary

results, published in ref. 25, that identified 5 and 25 minutes as

the optimal reporters of early phosphorylation activities. At

the end of the treatment cell lysates were collected using a

standard lysate procedure.25 Lysates from the two time points

were pooled together in a 1 : 1 ratio and the mixed lysate—

that corresponds to an ‘‘average early signaling response’’—

was measured using the Luminex xMAP technology. Mixing

cell lysates serves multiple purposes such as significant decrease

of experimental cost and improvement of data quality.30,51

In the 3rd day of the experimental procedure (follow-up

phosphoproteomic dataset) a multicombinatorial treatment of

cells is needed with the addition of inhibitors. In this case,

primary human hepatocytes were pre-treated with inhibitors

for 40 minutes. Two very selective kinase inhibitors were

chosen at concentrations sufficient to inhibit at least 95%

phosphorylation of the nominal target as determined by dose–

response assays in previous work.25 The following concentra-

tions were chosen: MEK/PD325901 (100 nM), PI3K/PI-103

(10 mM). After the pretreatment period, cells were stimulated

with a combination of stimuli selected from a library of 16

cytokines using the same concentrations and the same signals

were measured as in the first step.

Gaussian bimodal distribution for ligand selection and data

normalization

The Gaussian Mixture Distribution (GMD) was used at the

core of the ligand selection procedure. The measured phos-

phorylation values of each of the 14 signals were used to fit a

GMD consisting of 2 modes, the deactivated and the activated

mode. Subsequently for each measured phosphorylation value,

the probability distribution function (PDF) was evaluated

respective to both modes: if the PDF respective to the first

mode (deactivated) was greater than the PDF respective to the

second mode (activated) then the signal was considered to be

ON (otherwise OFF). This way the experimental dataset was

discretized to 0 and 1 (OFF and ON). Only stimuli that

activated at least one of the signals were included in the

follow-up steps of the proposed methodology. The Statistics

Toolbox of Matlab was used for this analysis and more

specifically the gmdistribution.fit() and pdf() functions.

A different normalization procedure was used for the com-

binatorial dataset, since the large number of experiments

performed with a strong activator and in the presence of

inhibitors led to very dispersed distributions that complicated

data discretization. The normalization scheme used is similar

to the one introduced in ref. 26. The fold change increase of

the signal (before and after stimulation) is evaluated and

passed through a hill function filter to be scaled from 0 to 1.

In more detail:

xk;mj jnormalized ¼

xkmj jt¼1xkmjjt¼0

� �n

pn þ xkmjjt¼1

xkmjjt¼0

� �n

where

xk,mj |normalized is the normalized measured value of species j

in experiment k,

xk,mj |t=0 is the unstimulated measured value of species j in

experiment k,

xk,mj |t=1 is the stimulated measured value of species j in

experiment k,

n is the hill coefficient, herein n = 4,

p is a user defined threshold representing the fold change

increase beyond which the signal is considered activated. In the

analysis presented herein p = 2, implying a 2-fold increase or

greater must take place for a signal to be considered activated.

Clustering

Clustering is performed using TM4 MeV (http://www.tm4.

org/mev/). Data are reconfigured into a 2-d matrix and

Dow

nloa

ded

by M

assa

chus

etts

Ins

titut

e of

Tec

hnol

ogy

on 1

9/04

/201

3 20

:47:

07.

Publ

ishe

d on

23

Mar

ch 2

012

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C2M

B05

482E

View Article Online

1582 Mol. BioSyst., 2012, 8, 1571–1584 This journal is c The Royal Society of Chemistry 2012

imported to MeV. Hierarchical clustering is performed respec-

tive to both axes according to the Euclidean distance and rows

and columns were reordered automatically. Other metrics

were also used, but offered little insight beyond what had

already been acquired using the Euclidean distance, therefore

are left out of the results section.

CellNOpt Preprocessing

Preprocessing of the generic pathway was performed using

CellNOpt50 (http://www.ebi.ac.uk/saezrodriguez/software.html)

to remove the non-observable and non-controllable part of it.

Node ‘‘A’’ is considered non-observable when its activation state

cannot be inferred based on the measured signals. This occurs

when there are no paths leading from ‘‘A’’ to any of the signals.

In CellNOpt the shortest path from ‘‘A’’ to each of the signals is

computed using Warshall’s algorithm, if such a path does not

exist, node ‘‘A’’ and all reactions leading from/to it are removed.

Node ‘‘B’’ is considered non-controllable, when its activation

state cannot be perturbed by any stimulus or inhibitor. This

occurs when there are no paths leading from any stimulus to this

node. Warshall’s algorithm is used to compute the shortest paths

from every stimulus to node ‘‘B’’, if such a path does not exist,

node B and all reactions leading from/to it are removed.

ILP formulation

Pathway optimization via an ILP formulation was first intro-

duced in ref. 51. Although the core of the formulation (i.e., the

inequality constraints) remains the same, the objective func-

tion is modified to account for the minimization of both the

measurement–prediction mismatch and the pathway size at the

same time, as in ref. 26 and 19. In ref. 51 the measurement–

prediction mismatch was minimized first resulting in a collec-

tion of solutions, where all the pathways scored the same

optimal-goodness of fit, and subsequently the pathway’s size

was minimized resulting in a solution that incorporates the

fewest reactions. The objective function used throughout the

paper is of the following form:

Xj:k

akj jxkj � xk;mj j þXi

biyi ð1Þ

where,

- j A {1,. . .,ns} is the set of species (nodes) included in the

pathway,

- i A {1,. . .,nr} is the set of reactions (edges) included in the

pathway,

- k A {1,. . .,ne} is the set of experiments,

- xkj is the predicted value of species j in the experiment k,

- xk,mj is the measured value of species j in experiment k,

- akj are user defined weights (for species j in experiment k)

- yi is the presence (or absence) of reaction i (assumes

Boolean values only)

- bi are user defined weights (for reaction i).

The first term of (1) corresponds to the measurement–pre-

diction mismatch, and its minimization guarantees the good-

ness of fit of the solution. The summation is performed only

over the measured species for each experiment. The second

term of (1) if bi > 0 minimizes the size of the pathway, else if

bi o 0 maximizes the size of the pathway.

The linear constraints that complete the formulation are:

zki r yi, i = 1,. . .,nr, k =1,. . .,ne. (2)

zki r xkj i = 1,. . .,nr, k =1,. . .,ne, j A Ri (3)

zki r 1 � xkj i = 1,. . .,nr, k =1,. . .,ne, j A li. (4)

zki � yi þXj2Ri

ðxkj � 1Þ�Xj2Iiðxkj Þ; i ¼ 1; . . . ; nr; k ¼ 1; . . . ; ne:

ð5Þ

xki Z zkj i = 1,. . .,nr, k =1,. . .,ne, j A Pi. (6)

xkj �X

i¼1;...;nr:j2Pi

zki ; j ¼ 1; . . . ; ns; k ¼ 1; . . . ; ne: ð7Þ

xkj = 0, k =1,. . .,ne, j A Mk,0 (8)

xkj = 1, k =1,. . .,ne, j A Mk,1 (9)

X A [0,1]ne�ns, y A {0,1}nr, Z A [0,1]ne�nr, (10)

where,

- zki corresponds to the activation or not (1 or 0 respectively)

of equation i in experiment k,

- Ri is the set of reactants in reaction i,

- Pi is the set of products in reaction i,

- Mk,1 is a set of species introduced in the system in

experiment k (imposed treatments),

- Mk,0 is a set of species excluded from the system in

experiment k.

Complete details of the ILP formulation can be found in

ref. 51. Note in particular that herein the variables xkj and zkiare relaxed to continuous. As demonstrated in ref. 27 this

results in an exact reformulation.

The resulting MILP is of significant size with 365 binary

variables and approximately 60000 continuous variables and

200000 constraints. Gurobi version 3.0.1 was used to solve this

problem to sufficiently tight tolerance (optcr = 0.10) in an

acceptable computational time of approximately 1.5 hours. Note

that the tractable computational time is in part achieved by the

relaxation of the integer variables xkj (species j present in experi-

ment k) and zkj (reaction i activated in experiment k) to continuous

(i.e., xkj A [0,1], zki A [0,1]) proposed in ref. 51. Note also that the

size of the MILP is drastically reduced by the preprocessing of the

signaling network performed by CellNetOptimizer.

Acknowledgements

We would like to thank Katerina Chairakaki for help with the

experimental procedure, Andreas Schuppert for help with the

experimental design, and Anastasios Papanastasiou for imple-

menting the semi-automated text mining procedure for identi-

fying optimal ligand concentrations. LGA and INM are

funded by the Marie Curie International Reintegration Grant

(MIRG-14-CT-2007-046531), and the Virtual Liver Network.

AM is grateful for funding through the Rockwell Inter-

national Career Development Professorship. Financial support

from the Deutsche Forschungsgemeinschaft (German Research

Association) through grant GSC 111 is gratefully acknowledged.

We would like to thank Gurobi Optimization for providing

the Gurobi solver free of charge.

Dow

nloa

ded

by M

assa

chus

etts

Ins

titut

e of

Tec

hnol

ogy

on 1

9/04

/201

3 20

:47:

07.

Publ

ishe

d on

23

Mar

ch 2

012

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C2M

B05

482E

View Article Online

This journal is c The Royal Society of Chemistry 2012 Mol. BioSyst., 2012, 8, 1571–1584 1583

References

1 K. Aggarwal and K. H. Lee, Briefings Funct. Genomics Proteomics,2003, 2, 175–184.

2 H. Kitano, Science, 2002, 295, 1662–1664.3 S. Boccaletti, V. Latora, Y. Moreno, M. Chavez and D. U. Hwang,Phys. Rep., 2006, 424, 175–308.

4 J. Saez-Rodriguez, L. G. Alexopoulos and G. Stolovitzky, Sci.Signaling, 2011, 4, 10.

5 L. Stromback and P. Lambrix, Bioinformatics, 2005, 21,4401–4407.

6 M. Hucka, A. Finney, H. M. Sauro, H. Bolouri, J. C. Doyle,H. Kitano, A. P. Arkin, B. J. Bornstein, D. Bray, A. Cornish-Bowden, A. A. Cuellar, S. Dronov, E. D. Gilles, M. Ginkel,V. Gor, Goryanin II, W. J. Hedley, T. C. Hodgman, J. H.Hofmeyr, P. J. Hunter, N. S. Juty, J. L. Kasberger, A. Kremling,U. Kummer, N. Le Novere, L. M. Loew, D. Lucio, P. Mendes,E. Minch, E. D. Mjolsness, Y. Nakayama, M. R. Nelson, P. F.Nielsen, T. Sakurada, J. C. Schaff, B. E. Shapiro, T. S. Shimizu,H. D. Spence, J. Stelling, K. Takahashi, M. Tomita, J. Wagner,J. Wang and S. Forum, Bioinformatics, 2003, 19, 524–531.

7 A. Morgat, E. Coissac, E. Coudert, K. B. Axelsen, G. Keller,A. Bairoch, A. Bridge, L. Bougueleret, I. Xenarios and A. Viari,Nucleic Acids Res., 2012, 40, D761–D769.

8 E. Demir, M. P. Cary, S. Paley, K. Fukuda, C. Lemer, I. Vastrik,G. Wu, P. D’Eustachio, C. Schaefer, J. Luciano, F. Schacherer,I. Martinez-Flores, Z. Hu, V. Jimenez-Jacinto, G. Joshi-Tope,K. Kandasamy, A. C. Lopez-Fuentes, H. Mi, E. Pichler,I. Rodchenkov, A. Splendiani, S. Tkachev, J. Zucker,G. Gopinath, H. Rajasimha, R. Ramakrishnan, I. Shah,M. Syed, N. Anwar, O. Babur, M. Blinov, E. Brauner,D. Corwin, S. Donaldson, F. Gibbons, R. Goldberg,P. Hornbeck, A. Luna, P. Murray-Rust, E. Neumann,O. Reubenacker, M. Samwald, M. van Iersel, S. Wimalaratne,K. Allen, B. Braun, M. Whirl-Carrillo, K. H. Cheung,K. Dahlquist, A. Finney, M. Gillespie, E. Glass, L. Gong, R. Haw,M. Honig, O. Hubaut, D. Kane, S. Krupa, M. Kutmon, J. Leonard,D. Marks, D. Merberg, V. Petri, A. Pico, D. Ravenscroft, L. Ren,N. Shah, M. Sunshine, R. Tang, R. Whaley, S. Letovksy,K. H. Buetow, A. Rzhetsky, V. Schachter, B. S. Sobral,U. Dogrusoz, S. McWeeney, M. Aladjem, E. Birney, J. Collado-Vides, S. Goto, M. Hucka, N. Le Novere, N. Maltsev, A. Pandey,P. Thomas, E. Wingender, P. D. Karp, C. Sander and G. D. Bader,Nat. Biotechnol., 2010, 28, 935–942.

9 P. Shannon, A. Markiel, O. Ozier, N. S. Baliga, J. T. Wang,D. Ramage, N. Amin, B. Schwikowski and T. Ideker, Genome Res.,2003, 13, 2498–2504.

10 S. Maere, K. Heymans and M. Kuiper, Bioinformatics, 2005, 21,3448–3449.

11 D. Smedley, S. Haider, B. Ballester, R. Holland, D. London,G. Thorisson and A. Kasprzyk, BMC Genomics, 2009, 10, 22.

12 A. Garg, K. Mohanram, G. De Micheli and I. Xenarios, MethodsMol. Biol. (Totowa, N. J.), 2012, 786, 397–443.

13 R. Apweiler, M. J. Martin, C. O’Donovan, M. Magrane, Y. Alam-Faruque, R. Antunes, D. Barrell, B. Bely, M. Bingley, D. Binns,L. Bower, P. Browne, W. M. Chan, E. Dimmer, R. Eberhardt,F. Fazzini, A. Fedotov, R. Foulger, J. Garavelli, L. G. Castro,R. Huntley, J. Jacobsen, M. Kleen, K. Laiho, D. Legge, Q. A. Lin,W. D. Liu, J. Luo, S. Orchard, S. Patient, K. Pichler, D. Poggioli,N. Pontikos, M. Pruess, S. Rosanoff, T. Sawford, H. Sehra,E. Turner, M. Corbett, M. Donnelly, P. van Rensburg,I. Xenarios, L. Bougueleret, A. Auchincloss, G. Argoud-Puy,K. Axelsen, A. Bairoch, D. Baratin, M. C. Blatter, B. Boeckmann,J. Bolleman, L. Bollondi, E. Boutet, S. B. Quintaje, L. Breuza,A. Bridge, E. deCastro, E. Coudert, I. Cusin, M. Doche,D. Dornevil, S. Duvaud, A. Estreicher, L. Famiglietti,M. Feuermann, S. Gehant, S. Ferro, E. Gasteiger, A. Gateau,V. Gerritsen, A. Gos, N. Gruaz-Gumowski, U. Hinz, C. Hulo,N. Hulo, J. James, S. Jimenez, F. Jungo, T. Kappler, G. Keller,V. Lara, P. Lemereier, D. Lieberherr, X. Martin, P. Masson,M. Moinat, A. Morgat, S. Paesano, I. Pedruzzi, S. Pilbout,S. Poux, M. Pozzato, N. Redaschi, C. Rivoire, B. Roechert,M. Schneider, C. Sigrist, K. Sonesson, S. Staehli, E. Stanley,A. Stutz, S. Sundaram, M. Tognolli, L. Verbregue, A. L. Veuthey,C. H. Wu, C. N. Arighi, L. Arminski, W. C. Barker, C. M. Chen,

Y. X. Chen, P. Dubey, H. Z. Huang, R. Mazumder, P. McGarvey,D. A. Natale, T. G. Natarajan, J. Nchoutmboube, N. V. Roberts,B. E. Suzek, U. Ugochukwu, C. R. Vinayaka, Q. H. Wang,Y. Q. Wang, L. S. Yeh, J. A. Zhang and U. Consortium, NucleicAcids Res., 2011, 39, D214–D219.

14 S. Kachalo, R. R. Zhang, E. Sontag, R. Albert and B. DasGupta,Bioinformatics, 2008, 24, 293–295.

15 J. P. Cobb, M. N. Mindrinos, C. Miller-Graziano, S. E. Calvano,H. V. Baker, W. Xiao, K. Laudanski, B. H. Brownstein,C. M. Elson, D. L. Hayden, D. N. Herndon, S. F. Lowry,R. V. Maier, D. A. Schoenfeld, L. L. Moldawer, R. W. Davis,R. G. Tompkins, P. Bankey, T. Billiar, D. Camp, I. Chaudry,B. Freeman, R. Gamelli, N. Gibran, B. Harbrecht, W. Heagy,D. Heimbach, J. Horton, J. Hunt, J. Lederer, J. Mannick,B. McKinley, J. Minei, E. Moore, F. Moore, R. Munford,A. Nathens, G. O’Keefe, G. Purdue, L. Rahme, D. Remick,M. Sailors, M. Shapiro, G. Silver, R. Smith, G. Stephanopoulos,G. Stormo, M. Toner, S. Warren, M. West, S. Wolfe andV. Young, Proc. Natl. Acad. Sci. U. S. A., 2005, 102, 4801–4806.

16 K. Basso, A. A. Margolin, G. Stolovitzky, U. Klein, R. Dalla-Faveraand A. Califano, Nat. Genet., 2005, 37, 382–390.

17 S. Bornholdt, Science, 2005, 310, 449–451.18 B. Schwanhausser, D. Busse, N. Li, G. Dittmar, J. Schuchhardt,

J. Wolf, W. Chen and M. Selbach, Nature, 2011, 473, 337–342.19 L. Anderson and J. Seilhamer, Electrophoresis, 1997, 18, 533–537.20 A. Pandey and M. Mann, Nature, 2000, 405, 837–846.21 L. G. Alexopoulos, J. Saez-Rodriguez and C. W. Espelin, in Drug

Efficacy, Safety, and Biologics Discovery: Emerging Technologiesand Tools, ed. S. Ekins and J. J. Xu, Wiley, 2009, pp. 29–52.

22 R. M. Ewing, P. Chu, F. Elisma, H. Li, P. Taylor, S. Climie,L. McBroom-Cerajewski, M. D. Robinson, L. O’Connor, M. Li,R. Taylor, M. Dharsee, Y. Ho, A. Heilbut, L. Moore, S. Zhang,O. Ornatsky, Y. V. Bukhman, M. Ethier, Y. Sheng, J. Vasilescu,M. Abu-Farha, J. P. Lambert, H. S. Duewel, Stewart II, B. Kuehl,K. Hogue, K. Colwill, K. Gladwish, B. Muskat, R. Kinach,S. L. Adams, M. F. Moran, G. B. Morin, T. Topaloglou andD. Figeys, Mol. Syst. Biol., 2007, 3, 89.

23 J. F. Rual, K. Venkatesan, T. Hao, T. Hirozane-Kishikawa,A. Dricot, N. Li, G. F. Berriz, F. D. Gibbons, M. Dreze,N. Ayivi-Guedehoussou, N. Klitgord, C. Simon, M. Boxem,S. Milstein, J. Rosenberg, D. S. Goldberg, L. V. Zhang,S. L. Wong, G. Franklin, S. Li, J. S. Albala, J. Lim,C. Fraughton, E. Llamosas, S. Cevik, C. Bex, P. Lamesch,R. S. Sikorski, J. Vandenhaute, H. Y. Zoghbi, A. Smolyar,S. Bosak, R. Sequerra, L. Doucette-Stamm, M. E. Cusick,D. E. Hill, F. P. Roth and M. Vidal, Nature, 2005, 437, 1173–1178.

24 J. V. Olsen, M. Vermeulen, A. Santamaria, C. Kumar,M. L. Miller, L. J. Jensen, F. Gnad, J. Cox, T. S. Jensen,E. A. Nigg, S. Brunak and M. Mann, Sci. Signaling, 2010, 3, ra3.

25 L. G. Alexopoulos, J. Saez-Rodriguez, B. D. Cosgrove,D. A. Lauffenburger and P. K. Sorger, Mol. Cell. Proteomics,2010, 9, 1849–1865.

26 R. J. Prill, D. Marbach, J. Saez-Rodriguez, P. K. Sorger,L. G. Alexopoulos, X. Xue, N. D. Clarke, G. Altan-Bonnet andG. Stolovitzky, PLoS One, 2010, 5, e9202.

27 B. B. Aldridge, J. M. Burke, D. A. Lauffenburger and P. K. Sorger,Nat. Cell Biol., 2006, 8, 1195–1203.

28 J. J. Hughey, T. K. Lee and M. W. Covert,Wiley Interdiscip. Rev.:Syst. Biol. Med., 2010, 2, 194–209.

29 M. K. Morris, J. Saez-Rodriguez, D. C. Clarke, P. K. Sorger andD. A. Lauffenburger, PLoS Comput. Biol., 2011, 7, e1001099.

30 I. Melas, A. Mitsos, D. Messinis, T. Weiss and L. Alexopoulos,BMC Syst. Biol., 2011, 5, 107.

31 K. Y. Yeung and W. L. Ruzzo, Bioinformatics, 2001, 17, 763–774.32 J. C. Liao, R. Boscolo, Y. L. Yang, L. M. Tran, C. Sabatti and

V. P. Roychowdhury, Proc. Natl. Acad. Sci. U. S. A., 2003, 100,15522–15527.

33 R. Albert, B. Dasgupta, R. Dondi, S. Kachalo, E. Sontag,A. Zelikovsky andK.Westbrooks, J. Comput. Biol., 2007, 14, 927–949.

34 A. Funahashi, Y. Matsuoka, A. Jouraku, M. Morohashi,N. Kikuchi and H. Kitano, Proc. IEEE, 2008, 96, 1254–1265.

35 B. B. Aldridge, J. Saez-Rodriguez, J. L. Muhlich, P. K. Sorger andD. A. Lauffenburger, PLoS Comput. Biol., 2009, 5, e1000340.

36 S. Klamt, J. Saez-Rodriguez and E. D. Gilles, BMC Syst. Biol.,2007, 8, 1–2.

Dow

nloa

ded

by M

assa

chus

etts

Ins

titut

e of

Tec

hnol

ogy

on 1

9/04

/201

3 20

:47:

07.

Publ

ishe

d on

23

Mar

ch 2

012

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C2M

B05

482E

View Article Online

1584 Mol. BioSyst., 2012, 8, 1571–1584 This journal is c The Royal Society of Chemistry 2012

37 J. Saez-Rodriguez, L. Simeoni, J. A. Lindquist, R. Hemenway,U. Bommhardt, B. Arndt, U. U. Haus, R. Weismantel, E. D. Gilles,S. Klamt and B. Schraven, PLoS Comput. Biol., 2007, 3, 1580–1590.

38 M. K. Morris, J. Saez-Rodriguez, P. K. Sorger and D. A.Lauffenburger, Biochemistry, 2010, 49, 3216–3224.

39 J. Thakar, M. Pilione, G. Kirimanjeswara, E. T. Harvill andR. Albert, PLoS Comput. Biol., 2007, 3, 1022–1039.

40 T. Helikar, J. Konvalina, J. Heidel and J. A. Rogers, Proc. Natl.Acad. Sci. U. S. A., 2008, 105, 1913–1918.

41 S. Bornholdt, J. R. Soc., Interface, 2008, 5(Suppl 1), S85–S94.42 L. Calzone, L. Tournier, S. Fourquet, D. Thieffry, B. Zhivotovsky,

E. Barillot and A. Zinovyev, PLoS Comput. Biol., 2010, 6, e1000702.43 C. Chaouiya, A. Naldi and D. Thieffry, Methods Mol. Biol.

(Totowa, N. J.), 2012, 804, 463–479.44 L. Mendoza and I. Xenarios, Theor. Biol. Med. Modell., 2006,

3, 13, DOI: 10.1186/1742-4682-3-13.45 S. Watterson, S. Marshall and P. Ghazal, Drug Discovery Today,

2008, 13, 447–456.46 K. Klemm and S. Bornholdt, Proc. Natl. Acad. Sci. U. S. A., 2005,

102, 18414–18419.47 A. Naldi, J. Carneiro, C. Chaouiya and D. Thieffry, PLoS Comput.

Biol., 2010, 6, e1000912.48 O. Sahin, H. Frohlich, C. Lobke, U. Korf, S. Burmester,

M. Majety, J. Mattern, I. Schupp, C. Chaouiya, D. Thieffry,A. Poustka, S. Wiemann, T. Beissbarth and D. Arlt, BMC Syst.Biol., 2009, 3, 1, DOI: 10.1186/1752-0509-3-1.

49 S. Uhlmann, H. Mannsperger, J. D. Zhang, E. A. Horvat,C. Schmidt, M. Kublbeck, F. Henjes, A. Ward, U. Tschulena,K. Zweig, U. Korf, S. Wiemann and O. Sahin, Mol. Syst. Biol.,2012, 8, 570.

50 J. Saez-Rodriguez, L. G. Alexopoulos, J. Epperlein, R. Samaga,D. A. Lauffenburger, S. Klamt and P. K. Sorger, Mol. Syst. Biol.,2009, 5, 331.

51 A. Mitsos, I. N. Melas, P. Siminelakis, A. D. Chairakaki,J. Saez-Rodriguez and L. G. Alexopoulos, PLoS Comput. Biol.,2009, 5, e1000591.

52 T. S. Weiss, S. Pahernik, I. Scheruebl, K.-W. Jauch andW. E. Thasler, J. Hepatology, 2003, 38, 476–482.

53 J. Saez-Rodriguez, A. Goldsipe, J. Muhlich, L. G. Alexopoulos,B. Millard, D. A. Lauffenburger and P. K. Sorger, Bioinformatics,2008, 24, 840–847.

54 A. Brooke, D. Kendrick and A. Meeraus, GAMS: User’s Guide,The Scientific Press, Redwood City, California, 1988.

55 K. Luckert, F. Gotschel, P. K. Sorger, A. Hecht, T. O. Joosand O. Potz, Mol. Cell. Proteomics, 2011, 10, M110.007377,Epub 4 March 2011.

56 B. Gao, W. I. Jeong and Z. G. Tian, Hepatology, 2008, 47,729–736.

57 F. Zhou, M. N. Ajuebor, P. L. Beck, T. Le, C. M. Hogaboam andM. G. Swain, Hepatology, 2005, 42, 372–380.

58 W. E. Thasler, T. S. Weiss, K. Schillhorn, P.-T. Stoll, B. Irrgangand K.-W. Jauch, Cell Tissue Banking, 2003, 4, 49–56.

Dow

nloa

ded

by M

assa

chus

etts

Ins

titut

e of

Tec

hnol

ogy

on 1

9/04

/201

3 20

:47:

07.

Publ

ishe

d on

23

Mar

ch 2

012

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C2M

B05

482E

View Article Online