Algorithmic and complexity results for decompositions of...

18
UNCORRECTED PROOF BIO 2594 1–18 BIO 2594 1–18 BioSystems xxx (2006) xxx–xxx Algorithmic and complexity results for decompositions of biological networks into monotone subsystems 3 4 Bhaskar DasGupta a, 1, , German Andres Enciso b, 2 , Eduardo Sontag c, 3 , Yi Zhang a, 1 5 a Department of Computer Science, University of Illinois at Chicago, Chicago, IL 60607, United States 6 b Mathematical Biosciences Institute, 250 Mathematics Building, 231 W 18th Avenue, Columbus, OH 43210, United States 7 c Department of Mathematics, Rutgers University, New Brunswick, NJ 08903, United States 8 Received 23 January 2006; received in revised form 3 August 2006; accepted 3 August 2006 9 Abstract 10 A useful approach to the mathematical analysis of large-scale biological networks is based upon their decompositions into mono- tone dynamical systems. This paper deals with two computational problems associated to finding decompositions which are optimal in an appropriate sense. In graph-theoretic language, the problems can be recast in terms of maximal sign-consistent subgraphs. The theoretical results include polynomial-time approximation algorithms as well as constant-ratio inapproximability results. One of the algorithms, which has a worst-case guarantee of 87.9% from optimality, is based on the semidefinite programming relaxation approach of Goemans–Williamson [Goemans, M., Williamson, D., 1995. Improved approximation algorithms for maximum cut and satisfiability problems using semidefinite programming. J. ACM 42 (6), 1115–1145]. The algorithm was implemented and tested on a Drosophila segmentation network and an Epidermal Growth Factor Receptor pathway model, and it was found to perform close to optimally. 11 12 13 14 15 16 17 18 19 © 2006 Elsevier Ireland Ltd. All rights reserved. 20 21 1. Introduction 22 In living cells, networks of proteins, RNA, DNA, 23 metabolites, and other species process environmental 24 signals, control internal events such as gene expres- 25 sion, and produce appropriate cellular responses. The 26 field of systems (molecular) biology is largely concerned 27 with the study of such networks, viewed as dynamical 28 systems. One approach to their mathematical analysis 29 Corresponding author. Tel.: +1 3123551319; fax: +1 3124130024. E-mail addresses: [email protected] (B. DasGupta), [email protected] (Y. Zhang), [email protected] (G.A. Enciso), [email protected] (E. Sontag). 1 Partly supported by NSF grants CCR-0296041, CCR-0206795, CCR-0208749 and IIS-0346973. 2 Work done while the author was with the Mathematics Depart- ment of Rutgers University and partly supported by NSF grant CCR- 0206789. 3 Partly supported by NSF grants EIA 0205116 and DMS-0504557. relies upon viewing them as made up of subsystems 30 whose behavior is simpler and easier to understand. Cou- 31 pled with appropriate interconnection rules, the hope is 32 that emergent properties of the complete system can be 33 deduced from the understanding of these subsystems. 34 Diagrammatically, we picture this as in Fig. 1, which 35 shows a full system as composed of four subsystems. 36 A particularly appealing class of candidates for “sim- 37 pler behaved” subsystems are monotone systems, as in 38 Hirsch (1985, 1983) and Smith (1995). Monotone sys- 39 tems are a class of dynamical systems for which patho- 40 logical behavior (“chaos”) is ruled out. Even though 41 they may have arbitrarily large dimensionality, mono- 42 tone systems behave in many ways like one-dimensional 43 systems. For instance, in monotone systems, bounded 44 trajectories generically converge to steady states, and 45 there are no stable oscillatory behaviors. More precisely, 46 see below, one must extend the notion of monotone sys- 47 tem so as to incorporate input and output channels, as 48 1 0303-2647/$ – see front matter © 2006 Elsevier Ireland Ltd. All rights reserved. 2 doi:10.1016/j.biosystems.2006.08.001 Please cite this article as: Bhaskar DasGupta et al., Algorithmic and complexity results for decompositions of biological networks into monotone subsystems, BioSystems (2006), doi:10.1016/j.biosystems.2006.08.001

Transcript of Algorithmic and complexity results for decompositions of...

Page 1: Algorithmic and complexity results for decompositions of ...dasgupta/resume/publ/papers/...UNCORRECTED PROOF BIO 2594 1–18 BIO25941–18 BioSystems xxx (2006) xxx–xxx Algorithmic

B

3

4

5

6

7

8

9

A10

tiToasat

11

12

13

14

15

16

17

18

19

©20

21

122

23

m24

s25

s26

fi27

w28

s29

y(

C

m0

1 02 d

D P

RO

OF

IO 2594 1–18

BioSystems xxx (2006) xxx–xxx

Algorithmic and complexity results for decompositions ofbiological networks into monotone subsystems

Bhaskar DasGuptaa,1,∗, German Andres Encisob,2, Eduardo Sontagc,3, Yi Zhanga,1

a Department of Computer Science, University of Illinois at Chicago, Chicago, IL 60607, United Statesb Mathematical Biosciences Institute, 250 Mathematics Building, 231 W 18th Avenue, Columbus, OH 43210, United States

c Department of Mathematics, Rutgers University, New Brunswick, NJ 08903, United States

Received 23 January 2006; received in revised form 3 August 2006; accepted 3 August 2006

bstract

A useful approach to the mathematical analysis of large-scale biological networks is based upon their decompositions into mono-one dynamical systems. This paper deals with two computational problems associated to finding decompositions which are optimaln an appropriate sense. In graph-theoretic language, the problems can be recast in terms of maximal sign-consistent subgraphs.he theoretical results include polynomial-time approximation algorithms as well as constant-ratio inapproximability results. Onef the algorithms, which has a worst-case guarantee of 87.9% from optimality, is based on the semidefinite programming relaxation

CT

pproach of Goemans–Williamson [Goemans, M., Williamson, D., 1995. Improved approximation algorithms for maximum cut andatisfiability problems using semidefinite programming. J. ACM 42 (6), 1115–1145]. The algorithm was implemented and tested on

Facto

30

31

32

33

34

E

Drosophila segmentation network and an Epidermal Growtho optimally.

2006 Elsevier Ireland Ltd. All rights reserved.

. Introduction

In living cells, networks of proteins, RNA, DNA,etabolites, and other species process environmental

ignals, control internal events such as gene expres-ion, and produce appropriate cellular responses. The

UN

CO

RReld of systems (molecular) biology is largely concerned

ith the study of such networks, viewed as dynamicalystems. One approach to their mathematical analysis

∗ Corresponding author. Tel.: +1 3123551319; fax: +1 3124130024.E-mail addresses: [email protected] (B. DasGupta),

[email protected] (Y. Zhang), [email protected]. Enciso), [email protected] (E. Sontag).1 Partly supported by NSF grants CCR-0296041, CCR-0206795,CR-0208749 and IIS-0346973.2 Work done while the author was with the Mathematics Depart-ent of Rutgers University and partly supported by NSF grant CCR-

206789.3 Partly supported by NSF grants EIA 0205116 and DMS-0504557.

35

36

37

38

39

40

41

42

43

44

303-2647/$ – see front matter © 2006 Elsevier Ireland Ltd. All rights reservoi:10.1016/j.biosystems.2006.08.001

Please cite this article as: Bhaskar DasGupta et al., Algorithmic and cinto monotone subsystems, BioSystems (2006), doi:10.1016/j.bios

Er Receptor pathway model, and it was found to perform close

relies upon viewing them as made up of subsystemswhose behavior is simpler and easier to understand. Cou-pled with appropriate interconnection rules, the hope isthat emergent properties of the complete system can bededuced from the understanding of these subsystems.Diagrammatically, we picture this as in Fig. 1, whichshows a full system as composed of four subsystems.

A particularly appealing class of candidates for “sim-pler behaved” subsystems are monotone systems, as inHirsch (1985, 1983) and Smith (1995). Monotone sys-tems are a class of dynamical systems for which patho-logical behavior (“chaos”) is ruled out. Even thoughthey may have arbitrarily large dimensionality, mono-tone systems behave in many ways like one-dimensionalsystems. For instance, in monotone systems, bounded

BIO 2594 1–18

trajectories generically converge to steady states, and 45

there are no stable oscillatory behaviors. More precisely, 46

see below, one must extend the notion of monotone sys- 47

tem so as to incorporate input and output channels, as 48

ed.

omplexity results for decompositions of biological networksystems.2006.08.001

Page 2: Algorithmic and complexity results for decompositions of ...dasgupta/resume/publ/papers/...UNCORRECTED PROOF BIO 2594 1–18 BIO25941–18 BioSystems xxx (2006) xxx–xxx Algorithmic

T

BIO 2594 1–18

2 B. DasGupta et al. / BioSystem

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

66

67

68

69

70

71

72

73

74

75

76

77

78

79

80

81

82

83

84

85

86

87

88

89

90

91

92

93

94

95

96

97

98

99

100

101

102

103

104

105

106

107

108

a given system. 109

C

Fig. 1. A system composed of four subsystems.

introduced and initially developed in Angeli and Sontag(2003); inputs and outputs are required so that intercon-nections like those shown in Fig. 1 can be defined.

Monotonicity is closely related, as explained later,to positive and feedback loops in systems. The topicof analyzing the behaviors of such feedback loops is along-standing one in biology in the context of regula-tion, metabolism, and development; a classical referencein that regard is the work (Monod and Jacob, 1961)of Monod and Jacob in 1961. See also, for example,Angeli et al. (2004), Angeli and Sontag (2004), Cinquinand Demongeot (2002), Lewis et al. (1977), Meinhardt(1978), Plathe et al. (1995), Remy et al. (2003), Snoussi(1998) and Thomas (1978).

An interconnection of monotone subsystems, that isto say, an entire system made up of monotone compo-nents, may or may not be monotone: “positive feedback”(in a sense that can be made precise) preserves mono-tonicity, while “negative feedback” destroys it. Thus,oscillators such as circadian rhythm generators requirenegative feedback loops in order for periodic orbits toarise, and hence are not themselves monotone systems,although they can be decomposed into monotone sub-

UN

CO

RR

Esystems (cf. Angeli and Sontag, 2004). A rich theory isbeginning to arise, characterizing the behavior of non-monotone interconnections. For example, Angeli andSontag (2003) shows how to preserve convergence to

Fig. 2. A consistent and an

Please cite this article as: Bhaskar DasGupta et al., Algorithmic andinto monotone subsystems, BioSystems (2006), doi:10.1016/j.bios

ED

PR

OO

F

s xxx (2006) xxx–xxx

equilibria; see also the follow-up papers (Angeli et al.,2004; Enciso et al., 2005; Enciso and Sontag, 2006;Gedeon and Sontag, 2005; De Leenheer et al., 2005).Even for monotone interconnections, the decomposi-tion approach is very useful, as it permits locating andcharacterizing the stability of steady states based uponinput/output behaviors of components, as described inAngeli and Sontag (2004); see also the follow-up papers(Angeli et al., 2004; Enciso and Sontag, 2005; De Leen-heer and Malisoff, 2006).

Moreover, a key point brought up in Sontag (2004,2005) is that new techniques for monotone systems inmany situations allow one to characterize the behaviorof an entire system, based upon the “qualitative” knowl-edge represented by general network topology and theinhibitory or activating character of interconnections,combined with only a relatively small amount of quan-titative data. The latter data may consist of steady-stateresponses of components (dose-response curves and soforth), and there is no need to know the precise formof dynamics or parameters such as kinetic constants inorder to obtain global stability conclusions.

In Section 2 of this paper, we briefly discuss mono-tonicity of systems described by ordinary differentialequations (the study of monotonicity can be extendedto partial differential equations, delay-differential equa-tions, and even more arbitrary dynamical systems, seee.g. Enciso and Sontag, 2006 in the context of mono-tone systems with inputs and outputs). We explain therehow the study of monotone systems, and more generallyof decompositions into monotone systems, relates to asign-consistency property for the graph which describeshow each state variable influences each other variable in

BIO 2594 1–18

inconsistent graph.

Generally, a graph, whose edges are labeled by “+” 110

or “−” signs (sometimes one writes +1, −1 instead of 111

+, −, or uses respectively activating “→” or inhibiting

complexity results for decompositions of biological networksystems.2006.08.001

Page 3: Algorithmic and complexity results for decompositions of ...dasgupta/resume/publ/papers/...UNCORRECTED PROOF BIO 2594 1–18 BIO25941–18 BioSystems xxx (2006) xxx–xxx Algorithmic

CT

BIO 2594 1–18

B. DasGupta et al. / BioSystem

“112

c113

s114

i115

e116

o117

d118

c119

e120

e121

o122

3123

r124

(125

126

t127

(128

(129

(130

(131

(132

n133

i134

d135

s136

t137

b138

c139

p140

t141

b142

t143

s144

d145

e146

T147

148

b149

p150

e151

m152

153

(154

g155

c156

a157

a158

159

160

161

162

163

164

165

166

167

168

169

170

171

172

173

174

175

176

177

178

179

180

181

182

183

184

185

186

187

188

189

190

191

192

193

194

195

196

197

198

199

200

201

202

203

204

205

206

NC

OR

RE

Fig. 3. Pulling-out inconsistent connections.

�” arrows as shown in Fig. 2), is said to be sign-onsistent if all paths between any two nodes have theame net sign, or equivalently, all closed loops have pos-tive parity, i.e. an even number, possibly 0, of negativedges. (For technical reasons, one ignores the directionf arrows, looking only at undirected graphs; see moreetails in Section 2.) Thus, the first graph in Fig. 2 isonsistent, but the second one, which differs in just onedge from the first one, is not (two paths with differ-nt parity are possible from node 1 to node 4, a directdd one as well as an even one transversing nodes 2 and). Self-loops, which in biochemical systems often rep-esent degradation terms, are ignored in this definition.We discuss this point further below.)

When applying decomposition theorems such ashose described in Angeli et al. (2004), Angeli et al.2004), Angeli and Sontag (2003, 2004), Enciso et al.2005), Enciso and Sontag (2005), Enciso and Sontag2006), Gedeon and Sontag (2005), De Leenheer et al.2005) and De Leenheer and Malisoff (2006), Sontag2004, 2005), it tends to be the case that the fewer theumber of interconnections among components, the eas-er it is to obtain useful conclusions. One may view aecomposition into interconnections of monotone sub-ystems as the “pulling out” of “inconsistent” connec-ions among monotone components, the original systemeing a “negative feedback” loop around an otherwiseonsistent system, as represented in Fig. 3. In this inter-retation, the number of interconnections among mono-one components corresponds to the number of variableseing fed-back. In addition, and independently from theheory developed in the above references, one mightpeculate that nature tends to favor systems that areecomposable into small monotone interconnections (orquivalently, have a small number of inconsistent paths).here are two reasons for this.

From a dynamical systems perspective, negative feed-ack loops, although required for homeostasis and foreriodic behavior, have potentially destabilizing effects,specially if there are signal propagation delays; thus,inimizing their number is desirable.Another advantage of consistency is as follows

Sontag, in preparation). Suppose that the nodes in the

Uraphs shown in Fig. 2 represent concentrations of ahemical species in a cell, such as receptors in a certainctivated state or transcription factors. Assume now thatperturbation instantaneously increases the value of the

Please cite this article as: Bhaskar DasGupta et al., Algorithmic and cinto monotone subsystems, BioSystems (2006), doi:10.1016/j.bios

ED

PR

OO

F

s xxx (2006) xxx–xxx 3

concentration of node 1. For the graph on the left, theinstantaneous effect on the other nodes is predictable:nodes 2 and 6 will increase, while nodes 3, 4, and 5will decrease. This unambiguous global effect holds trueregardless of the actual algebraic forms of reactions, val-ues of parameters such and kinetic constants, etc. Incontrast, consider the graph shown on the right. Nowthe net effect of an increase in node 1 is ambiguous. It isimpossible to know if node 4 will be repressed (becauseof the direct edge from 1 to 4) or activated (because ofthe indirect path). There is no way to resolve this ambi-guity unless equations and precise parameter values areassigned to the arrows. Since cells of the same type differin precise parameter values, due to varying concentra-tions of ATP, enzymes, and other chemicals, two cells ofthe same type may react in different ways to the same“stimulus” (increase in concentration of chemical 1).While such epigenetic diversity is sometimes desirable,it makes behavior less predictable. From an evolutionaryviewpoint, a “change in wiring” due to a mutation willhave an ambiguous effect, in this inconsistent network.

Of course, one should not expect large networks to beglobally consistent. However, if the number of inconsis-tencies in a biological interaction graph is small, it maywell be the case that the network is in fact consistentin a practical sense. For example, a gene regulatory net-work represents all potential effects among genes. Theseeffects are mediated by proteins which themselves mayneed to be “activated” in order to perform their func-tion, and this activation may, in turn, depend on certainextracellular ligands being present. Thus, depending onthe particular combination of external signals present,different subgraphs of the original graph describe thesystem under those conditions, and these graphs may beindividually consistent. For example, for the system inFig. 2, the edge from 1 to 2 may not be present under envi-ronmental conditions A, while the edge from 2 to 3 maynot be present under conditions B. Thus, under eitherconditions, A or B, the graph would be consistent, eventhough the entire network is not. See Sontag (in prepa-ration) for more discussion of these issues. In summary,consistency in biological networks may be desirable, andtherefore one might conjecture that true biological net-works tend to maximize it. Evidence that this is indeedthe case is provided by Ma’ayan et al. (in preparation),where the authors compare certain biological networksand appropriately randomized versions of them and showthat the original networks are closer to being consistent,

BIO 2594 1–18

when consistency is measured using a simple heuristic. 207

In the last section of this paper, we apply our algorithms 208

to perform a similar analysis, and once again derive the 209

conclusion that nature seems to favor consistency. 210

omplexity results for decompositions of biological networksystems.2006.08.001

Page 4: Algorithmic and complexity results for decompositions of ...dasgupta/resume/publ/papers/...UNCORRECTED PROOF BIO 2594 1–18 BIO25941–18 BioSystems xxx (2006) xxx–xxx Algorithmic

T

BIO 2594 1–18

4 B. DasGupta et al. / BioSystem

211

212

213

214

215

216

217

218

219

220

221

222

223

224

225

226

227

228

229

230

231

232

233

234

235

236

237

238

239

240

241

242

243

244

245

246

247

248

249

250

251

252

253

254

255

256

257

258

259

260

261

262

263

264

265

266

267

268

269

270

271

272

273

274

275

276

277

278

279

280

281

282

283

284

285

286

287

288

289

290

291

292

293

294

295

296

297

298

299

NC

OR

RE

C

Fig. 4. Dropping the diagonal edge gives consistency.

Thus, we are led to the subject of this paper, namelycomputing the smallest number of edges that have tobe removed so that there remains a consistent graph.For example, for the particular graph shown in Fig. 4the answer is that one edge (the diagonal positive one)suffices (in this case, the solution is unique: no singleother edge would suffice; in other problems, there maybe more than one optimizing solutions).

There has been other work dealing with efficientknock-out strategies in biochemical reaction networks,also formulated, as in this paper, as edge deletion prob-lems. As an example, we mention the recent paper(Klamt, 2006), which dealt with the question of iden-tifying a minimal set of reactions whose removal wouldblock the operation of a prespecified reaction. The prob-lem that we consider is completely different, however.

In this paper, we will study the computational com-plexity of the question of how many edges must beremoved in order to obtain consistency, and we pro-vide a relaxation-based polynomial-time approximationalgorithm guaranteed to solve the problem to about87.9% of the optimum solution, which is based onthe semidefinite programming relaxation approach ofGoemans–Williamson Goemans and Williamson (1995)(A variant of the problem is discussed as well.) We alsoobserve that it is not possible to have a polynomial-timealgorithm with performance too close to the optimal.While our emphasis is on theory, one of the algorithmswas implemented, and we show results of its applica-tion to a Drosophila segmentation network and to anEpidermal Growth Factor Receptor pathway model. Itturns out that, when applying the algorithm, often thesolution is much closer to optimal than the worst-caseguarantee of 87.9%, and indeed often gives an optimalsolution.

The remainder of this paper is organized as follows.Section 2 briefly discusses monotonicity. The discussionis self-contained for the purposes of this paper, and ref-erences are given to the dynamical systems results thatmotivate the problem studied here. The connection to

Uconsistency is also explained there. Section 3 discussesthe associated graph-theoretic problems and notions ofapproximability used in the paper, leading to the state-ment of our main theoretical results in Section 4, which

Please cite this article as: Bhaskar DasGupta et al., Algorithmic andinto monotone subsystems, BioSystems (2006), doi:10.1016/j.bios

ED

PR

OO

F

s xxx (2006) xxx–xxx

are proved in Section 5. Section 6 contains the men-tioned examples of application of the algorithm. Finally,in Section 6.3 we consider a yeast gene regulatory net-work and various randomized versions of it, concludingthat the original network is far closer to consistent thanmay be expected from chance alone. Several technicalproofs are separately provided in Appendix A.

2. Monotone systems and consistency

We will illustrate the motivation for the problem stud-ied here using systems of ordinary differential equations

x = F (x) (1)

(the dot indicates time derivative, and x = x(t) is a vec-tor), although the discussion applies as well to moregeneral types of dynamical systems such as delay-differential systems or certain systems of reaction-diffusion partial differential equations. In applicationsto biological networks, the component xi(t) of the vec-tor x = x(t) indicates the concentration of the ith speciesin the model at time t.

We will restrict attention to models in which the directeffect that one given variable in the model has overanother is unambiguous, in the sense that it is alwaysinhibitory or always promoting. Thus, if protein A bindsto the promoter region of gene B, we assume that it doesso either to prevent the transcription of the gene or tofacilitate it, no matter what are the respective concen-trations. Mathematically, what we are saying is that werequire that for every i, j = 1, . . . , n, i �= j, the partialderivative ∂Fi/∂xj be either ≥ 0 at all states or ≤ 0 at allstates.

Let us briefly discuss this non-ambiguity assump-tion. First of all, we remark that this assumption doesnot prevent protein A from having an indirect influ-ence, through other molecules, perhaps dimmers of Aitself, that can ultimately lead to the opposite effecton gene B from that of a direct connection. Indeed,this is the whole point of studying graph consistency.Second, in biomolecular networks, ambiguous signs inJacobians often represent heterogeneous mechanisms.For example, take the case where protein A enhances thetranscription rate of gene B only if it is present at low con-centrations, but represses B if its concentration is largerthan some threshold. A careful study of the chemicalmechanism often reveals the existence of an interme-diate form (perhaps a homodimer) that is responsible

BIO 2594 1–18

for this ambiguous effect. (Mathematically, an example 300

is a rate of transcription k1a − k2a2, where a denotes 301

the concentration of A.) Introducing a new species into 302

the model (mathematically, an additional state variable 303

complexity results for decompositions of biological networksystems.2006.08.001

Page 5: Algorithmic and complexity results for decompositions of ...dasgupta/resume/publ/papers/...UNCORRECTED PROOF BIO 2594 1–18 BIO25941–18 BioSystems xxx (2006) xxx–xxx Algorithmic

CT

B

oSystem

r304

p305

o306

c307

w308

r309

F310

d311

e312

o313

l314

J315

I316

i317

o318

g319

f320

v321

s322

t323

w324

t325

t326

327

(328

y329

a330

r331

o332

w333

i334

n335

s336

(337

p338

i339

340

o341

l342

f343

i344

o345

≤346

1347

e348

l349

p350

S351

s352

L353

s354

355

356

357

358

359

360

361

362

363

364

365

366

367

368

369

370

371

372

373

374

375

376

377

378

379

380

381

382

383

384

385

386

387

388

389

390

391

392

393

394

395

396

397

398

NC

OR

RE

IO 2594 1–18

B. DasGupta et al. / Bi

epresenting this intermediate form) reduces one to theroblem in which Jacobian entries are unambiguous. (Inur example, we would write the rate as k1a − k2c, whereis the concentration of the dimer. In addition, thereould be a new equation such as dc/dt = k3a

2 − k4c

epresenting formation of the dimer and its degradation.)inally, we note that small-scale negative loops are abun-ant in nature. Self-loops or “auto repression” are anxtreme example of these, and appear as a consequencef degradation and other effects. Regarding such self-oops, observe that the requirement of a fixed sign foracobian entries is not imposed on diagonal elements.n fact, these elements play no role in the graph to bentroduced next, nor on monotonicity—the propertiesf monotone systems are not affected by them. Moreenerally, it is often the case that small loops representast dynamics which may be collapsed into a self-loopsia time-scale decomposition (singular perturbations or,pecifically for enzymes, “quasi-steady state approxima-ions”) and hence may be viewed and diagonal termshich may be safely ignored. This is a modeling ques-

ion, to be settled before the algorithms studied here areo be applied.

Given any partial order ≤ defined on Rn, a system1) is said to be monotone with respect to ≤ if x0 ≤0 implies x(t) ≤ y(t) for every t ≥ 0. Here x(t), y(t)re the solutions of (1) with initial conditions x0, y0,espectively. Of course, whether a system is monotoner not depends on the partial order being considered, bute one says simply that a system is monotone if the order

s clear from the context. Monotonicity with respect toontrivial orders rules out chaotic attractors and eventable periodic orbits; see Hirsch (1985, 1983), Smith1995), and is, as discussed in the introduction, a usefulroperty for components when analyzing larger systemsn terms of subsystems.

A useful way to define partial orders in Rn, and thenly one to be further considered in this paper, is as fol-ows. Given a tuple s = (s1, . . . , sn), where si ∈ {1, −1}or every i, we say that x ≤s y if sixi ≤ siyi for every. For instance, the “cooperative order” is the orthantrder ≤s generated by s = (1, . . . , 1). This is the order

defined by x ≤ y if and only if xi ≤ yi for all i =, . . . , n. It is not difficult to verify if a system is coop-rative with respect to an orthant order; the followingemma, known as “Kamke’s condition,” is not hard torove, see Smith (1995) for details (also Angeli andontag, 2003 in the more general context of monotone

Uystems with input and output channels).

emma 1. Consider an orthant order ≤s generated by= (s1, . . . , sn). A system (1) is monotone with respect

Please cite this article as: Bhaskar DasGupta et al., Algorithmic and cinto monotone subsystems, BioSystems (2006), doi:10.1016/j.bios

ED

PR

OO

F

s xxx (2006) xxx–xxx 5

to ≤s if and only if

sisj∂Fj

∂xi

≥ 0, i, j = 1, . . . , n, i �= j. (2)

To provide intuition, let us sketch the sufficiency partof the proof for the special case of the cooperativeorder. Suppose by contradiction that the system is notmonotone, and that therefore there is a pair of ini-tial conditions x0 ≤ y0 whose solutions x(t), y(t) ceaseto satisfy x(t) ≤ y(t) at some point. This implies thatat a certain critical moment in time t, there is somecoordinate i so that xi(t−) < yi(t−) but xi(t+) > yi(t+).(This argument is not entirely accurate, but it givesthe flavor of the proof.) Thus xi(t) = yi(t) for some iand the derivative with respect to time of xi is largerthan that of yi at time t, meaning that that Fi(x) >

Fi(y), where x = xi(t) and y = yi(t). However, thiscannot happen if Fi is increasing on all the variablesxj except possibly xi, so that x ≤ y, xi = yi impliesFi(x) ≤ Fi(y). An equivalent way to phrase this con-dition is by ask that ∂Fi/∂xj ≥ 0 at all states for everyi, j, i �= j, which is the Kamke condition for the specialcase of the cooperative order. The name of the orderarises because in a monotone system with respect to thatorder each species promotes or “cooperates” with eachother.

A rephrasing of this characterization of monotonicitywith respect to orthant orders can be given by looking atthe signed digraph G associated to (1). We define thevertex set V (G) and the edge set E(G) of G as fol-lows. Let V (G) = {1, . . . , n}, and given vertices i, j,let (i, j) ∈ E(G) and fE(i, j) = 1 if both ∂Fj/∂xi ≥ 0and the strict inequality holds at least at one state.Similarly let (i, j) ∈ E(G) and fE(i, j) = −1 if both∂Fj/∂xi ≤ 0 and the strict inequality holds at least at onestate. Finally, let (i, j) �∈ E(G) if ∂Fj/∂xi ≡ 0. Recallthat we are assuming that one of the three cases musthold.

Now we can define an orthant cone using any func-tion fV : V (G) → {−1, 1}, by letting x ≤fV y if andonly if fV (i)xi ≤ fV (i)yi for all i. Given fV , we definethe consistency function g : E(G) → {true, false} byg(i, j) = fV (i)fV (j)fE(i, j). Then, the following analogof Lemma 1 holds.

Lemma 2. Consider a system (1) and an orthant cone≤fV . Then (1) is monotone with respect to ≤fV if and

BIO 2594 1–18

only if g(i, j) ≡ 1 on E(G). 399

Proof. Let si = fV (i), i = 1, . . . , n. Note that 400

sisj∂fi/∂xj = 0 if (i, j) �∈ E(G). For (i, j) ∈ E(G), it 401

holds that sisj∂fi/∂xj ≥ 0 if and only if sisjfE(i, j) = 1, 402

omplexity results for decompositions of biological networksystems.2006.08.001

Page 6: Algorithmic and complexity results for decompositions of ...dasgupta/resume/publ/papers/...UNCORRECTED PROOF BIO 2594 1–18 BIO25941–18 BioSystems xxx (2006) xxx–xxx Algorithmic

T

oSystem

403

404

405

406

407

408

409

410

411

412

413

414

415

416

417

418

419

420

421

422

423

424

425

426

427

428

429

430

431

432

433

434

435

436

437

438

439

440

441

442

443

444

445

446

447

448

449

450

451

452

453

454

455

456

457

458

459

460

461

462

463

464

465

466

467

468

469

470

471

472

473

474

475

476

477

478

479

480

481

482

483

484

485

486

487

488

489

490

491

492

NC

OR

RE

C

BIO 2594 1–18

6 B. DasGupta et al. / Bi

that is, if and only if g(i, j) = 1. The result follows fromLemma 1. �

For the next lemma, let the parity of a chain in G be theproduct of the signs (+1, −1) of its individual edges. Wewill consider in the next result closed undirected chains,that is, sequences xi1 , . . . , xir such that xi1 = xir , andsuch that for every λ = 1, . . . , r − 1 either (xiλ, xiλ+1 ) ∈E(G) or (xiλ+1 , xiλ ) ∈ E(G).

The following lemma (see DeAngelis et al., 1986 aswell as Smith, 1988, page 101) is analogous to the factfrom vector calculus that path integrals of a vector fieldare independent of the particular path of integration ifand only if there exists a potential function. Since theresult is key to the formulation of the problem beingconsidered, we provide a simple and self-contained proofin Appendix A.

Lemma 3. Consider a dynamical system (1) with asso-ciated directed graph G. Then (1) is monotone withrespect to some orthant order if and only if all closedundirected chains of G have parity 1.

2.1. Systems with inputs and outputs

As we discussed in the introduction, a usefulapproach to the analysis of biological networks consistsof decomposing a given system into an interconnectionof monotone subsystems. The formulation of the notionof interconnection requires subsystems to be endowedwith “input and output channels” through which infor-mation is to be exchanged. In order to address this weconsider controlled dynamical systems (Sontag, 1990)which are systems with an additional parameter u ∈ Rm

and which have the form

x = g(x, u). (3)

The values of u over time are specified by means ofa function t → u(t) ∈ Rm, t ≥ 0, called an input orcontrol. Thus each input defines a time-dependentdynamical system in the usual sense. To system (3)there is associated a feedback function h : Rn → R

m,which is usually used to create the closed loop systemx = g(x, h(x)). Finally, ifRn,Rm are ordered by orthantorders ≤fV , ≤q respectively, we say that the system ismonotone if it satisfies (2) for every u, and also

q f (j)∂gj ≥ 0, for every k, j (4)

Uk V∂uk

(see also Angeli and Sontag, 2003.) As an example, letus consider the following biological model of testos-terone dynamics (Enciso and Sontag, 2004; Murray and

Please cite this article as: Bhaskar DasGupta et al., Algorithmic andinto monotone subsystems, BioSystems (2006), doi:10.1016/j.bios

ED

PR

OO

F

s xxx (2006) xxx–xxx

Mathematical Biology, 2002):

x1 = A

K + x3− b1x1, x2 = c1x1 − b2x2,

x3 = c2x2 − b3x3. (5)

Drawing the digraph of this system, it is easy to see thatit is not monotone with respect to any orthant order,as follows by application of Lemma 3. On the otherhand, replacing x3 in the first equation by u, we obtaina system that is monotone with respect to the orders≤(1,1,1), ≤(−1) for state and input respectively. Definingh(x) = x3, the closed loop system of this controlledsystem is none other than (5). The paper (Enciso andSontag, 2004) shows how, using this decompositiontogether with the “small gain theorem” from monotoneinput/output theory (Angeli and Sontag, 2003) leadsone to a proof that the system does not have oscillatorybehavior, even under arbitrary delays in the feedbackloop, contrary to the assertion made in Murray andMathematical Biology (2002).

We can carry out this procedure on an arbitrary sys-tem (1) with a directed graph G, as follows: given aset E of edges in G, enumerate the edges in EC as(i1, j1), . . . , (im, jm). For every k = 1, . . . , m, replaceall appearances of xik in the function Fjk

by the vari-able uk, to form the function g(x, u). Define h(x) =(xi1 , . . . , xim ). It is easy to see that this controlled system(3) has closed loop (1).

Note that the controlled system (3) generated by theset E as above has, as associated digraph, the sub-digraphof G generated by E. This is because for every k, one has∂gjk

(x, u)/∂xik ≡ 0, i.e., the edge from ik to jk has been“erased”.

Denote by G the underlying undirected graph of adirected graph G obtained by ignoring the directions ofthe edges. Given a set E ⊆ V (G) of vertices in a (directedor undirected) graph G, denote by G(E) the undirectedsubgraph of G generated by E. The edges of both G andG(E) are labeled with ±1 using the labels in the edgesof G, whenever appropriate. Let E be called consistent ifG(E) has no closed chains with parity −1. Note that thisis equivalent to the existence of fV such that g ≡ 1 on E,by Lemma 4 applied to the open loop system (3). If E isconsistent, then the associated system (3) itself can alsobe shown to be monotone: to verify condition (4), sim-ply define each qk so that (4) is satisfied for k, jk. Since∂g /∂u = ∂F /∂x �≡ 0, this choice is in fact unam-

BIO 2594 1–18

jk k jk ik

biguous. Conversely, if (3) is monotone with respect to 493

the orthant orders ≤fV , ≤q, then in particular it is mono- 494

tone for every fixed constant u, so that E is consistent by 495

Lemma 3. We thus have the following result. 496

complexity results for decompositions of biological networksystems.2006.08.001

Page 7: Algorithmic and complexity results for decompositions of ...dasgupta/resume/publ/papers/...UNCORRECTED PROOF BIO 2594 1–18 BIO25941–18 BioSystems xxx (2006) xxx–xxx Algorithmic

CT

B

oSystem

L497

T498

c499

o500

3501

502

a503

t504

o505

i506

s507

‖508

p509

P510

i511

a512

t513

a514

f515

t516

e517

T518

c519

P520

a521

a522

s523

f524

D525

t526

t527

P528

s529

t530

s531

x532

A533

e534

3535

“536

|537

a538

|539

v540

F541

o542

a543

544

545

546

547

548

549

550

551

552

553

554

555

556

557

558

559

560

561

562

563

564

565

566

567

568

569

570

571

572

573

574

575

576

577

578

579

580

581

582

583

584

585

586

587

588

NC

OR

RE

IO 2594 1–18

B. DasGupta et al. / Bi

emma 4. Let E be a set of edges of the digraph G.hen E is consistent if and only if the correspondingontrolled system (3) is monotone with respect to somerthant orders.

. Statement of problem

A natural problem is therefore the following. Givendynamical system (1) that admits a digraph G, use

he procedure above to decompose it as the closed loopf a monotone controlled system (3), while minimiz-ng the number ‖EC‖ of inputs. Equivalently, find fV

uch that P(E+) = ‖E+‖ is maximized and P(E−) =E−‖ = ‖EC+‖ minimized. This produces the followingroblem formulation.

roblem 1 (Undirected labeling problem (ULP)). Annstance of this problem is (G, h), where G = (V, E) isn undirected graph and h : E �→ {0, 1}. A valid solu-ion is a vertex labeling function f : V → {0, 1}. Definen edge {u, v} ∈ E to be consistent iff h(u, v) ≡ (f (u) +(v)) (mod 2). The objective is then to find a valid solu-

ion maximizing |F | where F is the set of consistentdges.

hat ULP is a correct formulation for our problem isonfirmed by the following easy equivalence.

roposition 1. Consider an instance (G, h) of ULP withn optimal solution having x consistent edges given byvertex labeling function f. Let D be a set of edges of

mallest cardinality that have to be removed such thator the remaining graph, that is the graph G′ = (V, E \

) with the same vertex set V but an edge set E \ D,here exists a vertex labeling function f ′ : V → {0, 1}hat makes every edge consistent. Then, x = |E| − |D|.roof. Since f produces a solution of ULP with x con-istent edges, exactly |E| − x edges are inconsistent,hus |D| ≤ |E| − x, that is, x ≤ |E| − |D|. Conversely,ince there is a solution with |E| − |D| consistent edges,≥ |E| − |D|. �

special case of ULP, namely when h(e) = 1 for all∈ E, is the MAX-CUT problem (defined in Section.1). Moreover, ULP can be posed as a special type ofconstraint satisfaction problem” as follows. We haveE| linear equations over GF (2), one equation per edgend each equation involving exactly two variables, overV | Boolean variables. The goal is to assign values to the

Uariables to satisfy the maximum number of equations.or algorithms and lower-bound results for general casesf these types of problems, such as when the equationsre over GF (p) for an arbitrary prime p > 2, when there

Please cite this article as: Bhaskar DasGupta et al., Algorithmic and cinto monotone subsystems, BioSystems (2006), doi:10.1016/j.bios

ED

PR

OO

F

s xxx (2006) xxx–xxx 7

are an arbitrary number of variables per equation or whenthe goal is to minimize the number of unsatisfied equa-tions, see references such as Amaldi and Kann (1996),Berman and Karpinski (2001), Creignou et al. (2001) andHastad and Venkatesh (2002) and the references therein.

Another interpretation (Sontag, in preparation) ofULP is in statistical mechanics terms. Let us label edgesby “±1” instead of {0, 1}, denoting by wuv = (−1)h(u,v)

the edge parities, now called “interaction energies.” Sim-ilarly, let us consider ±1-valued vertex labeling func-tions, now called (magnetic) “spin configurations,” σ :V → {−1, +1}, σ(v) = (−1)f (v). An edge {u, v} is con-sistent provided that wuvσuσj = 1. A graph with ±1weights is called an Ising spin-glass model in statisticalphysics. A “non-frustrated” spin-glass model is one forwhich there is a spin configuration for which every edgeis consistent (Barahona, 1982; Cipra, 2000; De Simoneet al., 1995; Istrail, 2000). This is the same as a consis-tent graph in our sense. Moreover, a spin configurationthat maximizes the number of consistent edges is one forwhich the “free energy” (with no exterior magnetic field):

−∑ij

wuvσuσv

is minimized, a “ground state”. (When h(e) = 1 orequivalently we = −1 for all edges, one has whatis called the “anti-ferromagnetic case”.) Thus, ourproblem amounts to finding ground states.

Given orthant orders ≤fV and ≤q for Rn and Rm

respectively, we say that a feedback function h is positiveif x ≤fV y implies h(x) ≤q h(y), and that it is negativeif x ≤fV y implies h(x) ≥q h(y). It can be shown thatthe closed loop of a monotone system with a positivefeedback function is actually itself monotone, so that nosystem can be produced in this way that was not mono-tone already. But if h is a negative feedback function, thenseveral results become available which use the methodsof monotone systems for systems that are not monotone,see Angeli and Sontag (2003), Enciso and Sontag (2004)and Enciso and Sontag (2006). For the following result,let (C, ⊆) be the class of consistent subsets of E(G),ordered under inclusion.

Proposition 2. Let E be a consistent set. Then E ismaximal in (C, ⊆) if and only if h is a negative feedbackfunction for every fV such that g ≡ 1 on E.

Proof. Suppose that E is maximal, and let fV besuch that g ≡ 1 on E. Given any edge (i , j ) ∈ EC, it

BIO 2594 1–18

k k

holds that g(ik, jk) = −1. Otherwise one could extend 589

E by adding (ik, jk), thus violating maximality. That 590

is, fV (ik)fV (jk)fE(ik, jk) = −1. By monotonicity, it 591

holds that qkfV (jk)∂gjk/∂uk ≥ 0, and since ∂gjk

/∂uk = 592

omplexity results for decompositions of biological networksystems.2006.08.001

Page 8: Algorithmic and complexity results for decompositions of ...dasgupta/resume/publ/papers/...UNCORRECTED PROOF BIO 2594 1–18 BIO25941–18 BioSystems xxx (2006) xxx–xxx Algorithmic

T

oSystem

593

594

595

596

597

598

599

600

601

602

603

604

605

606

607

608

609

610

611

612

613

614

615

616

617

618

619

620

621

622

623

624

625

626

627

628

629

630

631

632

633

634

635

636

637

638

639

640

641

642

643

644

645

646

647

648

649

650

651

652

653

654

655

656

657

658

659

660

661

662

663

664

665

666

667

668

669

670

671

672

673

674

675

676

677

678

679

680

681

682

683

684

685

686

687

688

NC

OR

RE

C

BIO 2594 1–18

8 B. DasGupta et al. / Bi

∂Fjk/∂xik , it follows necessarily that

qkfV (jk)fE(ik, jk) = 1.

Therefore it must hold that qk = −fV (ik) for each k,which implies that h is a negative feedback function.

Conversely, if fV is such that g ≡ 1 on E and h is anegative feedback function, then qk = −fV (ik). By thesame argument as above, qkfV (jk)fE(ik, jk) = 1 for allk by monotonicity. Therefore g ≡ −1 on EC. Repeatingthis for all admissible fV , maximality follows. �There is a second, slightly more sophisticated way ofwriting a system (1) as the feedback loop of a system (3)using an arbitrary set of edges E. Given any such E,define S(Ec) = {i|there is some jsuch that (i, j) ∈ Ec}.Now enumerate S(Ec) as {i1, . . . , im}, and for each klabel the set {j|(ik, j) ∈ Ec} as jk1, jk2, . . .. Then foreach k, l, one can replace each appearance of xik inFjkl

by uk, to form the function g(x, u). Then one letsh(x) = (xi1 , . . . , xim ) as above. The closed loop of thissystem (3) is also (1) as before but with the advantage thatthere are |S(Ec)| inputs, and of course |S(Ec)| ≤ |Ec|.

If E is a consistent and maximal set, then one canmake (3) into a monotone system as follows. By let-ting fV be such that g ≡ 1 on E, we define the order≤fV on Rn. For every ik, jkl such that (ik, jkl) ∈ EC,it must hold that fV (ik)fV (jkl)fE(ik, jkl) = −1. Other-wise E ∪ {(ik, jkl)} would be consistent, thus violatingmaximality. By choosing qk = −fV (ik), Eq. (4) is there-fore satisfied. See the proof of Proposition 2. Conversely,if the system generated by E using this second algorithmis monotone with respect to orthant orders, and if h is anegative function, then it is easy to verify that E must beboth consistent and maximal.

Thus the problem of finding E consistent and suchthat P(E−) = ‖S(E−)‖ = ‖S(EC)‖ is smallest, whenrestricted to those sets that are maximal and consistent(this does not change the minimum ‖S(EC)‖), is equiv-alent to the following problem: decompose (1) into thenegative feedback loop of an orthant monotone controlsystem, using the second algorithm above, and using asfew inputs as possible. This produces the following prob-lem formulation.

Problem 2 (Directed labeling problem (DLP)). Aninstance of this problem is (G, h) where G = (V, E) isa directed graph and h : E → {0, 1}. A valid solutionis a vertex labeling function f : V → {0, 1}. Define anedge (u, v) ∈ E to be consistent iff h(u, v) ≡ (f (u) +

Uf (v)) (mod 2). The objective is then to find a validsolution minimizing |g(E − F )| where g(C) = {u ∈ V |∃y ∈ V, (u, y) ∈ C} for any C ⊆ E and F is the set ofconsistent edges.

Please cite this article as: Bhaskar DasGupta et al., Algorithmic andinto monotone subsystems, BioSystems (2006), doi:10.1016/j.bios

ED

PR

OO

F

s xxx (2006) xxx–xxx

3.1. Summary of key concepts and results inapproximation algorithms

For anyγ ≥ 1 (resp.γ ≤ 1), aγ-approximate solution(or simply an γ-approximation) of a minimization (resp.,maximization) problem is a solution with an objectivevalue no larger than γ times (resp., no smaller thatγ times) the value of the optimum, and an algorithmachieving such a solution is said to have an approxima-tion ratio of γ .

In Papadimitriou and Yannakakis (1991) Papadim-itriou and Yannakakis defined the class of MAX-SNPoptimization problems and a special approximation-preserving reduction, the so-called L-reduction, that canbe used to show MAX-SNP-hardness of an optimizationproblem. The version of the L-reduction that we providebelow is a slightly modified but equivalent version thatappeared in Berman and Schnitger (1992).

Definition 1. Berman and Schnitger (1992),Papadimitriou and Yannakakis (1991) Given two opti-mization problems Π and Π ′, we say that Π L-reduces toΠ ′ if there are three polynomial-time procedures T1,T2,T3 and two constants a and b > 0 such that the followingtwo conditions are satisfied: (1) For any instance I of Π,algorithm T1 produces an instance I ′ = f (I) of Π ′ gen-erated from T1 such that the optima of I and I ′, OPT(I)and OPT(I ′), denoted by respectively, satisfy OPT(I ′) ≤a · OPT(I). (2) For any solution of I ′ with cost c′, algo-rithm T2 produces another solution with a cost c′′ noworse than c′, and algorithm T3 produces a solution ofI of Π with cost c (possibly from the solution producedby T2) satisfying |c − OPT(I)| ≤ b · ∣∣c′′ − OPT(I ′)

∣∣.An optimization problem is MAX-SNP-hard if any prob-lem in MAX-SNP L-reduces to that problem. The impor-tance of proving MAX-SNP-hardness results comesfrom a result proved by Arora et al. Arora et al. (1998)which shows that, assuming P �=NP, for every MAX-SNP-hard minimization (resp., maximization) problemthere exists a constant ε > 0 such that no polynomialtime algorithm can achieve an approximation ratio bet-ter than 1 + ε (resp., better than 1 − ε).

A special case of the ULP problem, namely whenh(e) = 1 for all e ∈ E, is the well-known MAX-CUTproblem. An instance of this problem is an undirectedgraph G = (V, E). A valid solution is a set S ⊆ V . Theobjective is to find a valid solution that maximizes thenumber of edges {u, v} ∈ E such that |{u, v} ∩ S| = 1.

BIO 2594 1–18

The MAX-CUT problem is known to be MAX-SNP- 689

hard. For further details on these topics, the reader is 690

referred to the excellent book by Vazirani (Vazirani, 691

2001). 692

complexity results for decompositions of biological networksystems.2006.08.001

Page 9: Algorithmic and complexity results for decompositions of ...dasgupta/resume/publ/papers/...UNCORRECTED PROOF BIO 2594 1–18 BIO25941–18 BioSystems xxx (2006) xxx–xxx Algorithmic

CT

B

oSystem

693

f694

a695

S696

d697

w698

n699

G700

r701

4702

703

T704

(705

706

707

708

(709

710

711

712

713

(714

715

716

717

O718

a719

n720

2721

G722

R723

U724

c725

t726

t727

e728

i729

s730

a731

s732

1733

i734

s735

e736

i737

t738

t739

i740

741

742

743

744

745

746

747

748

749

750

751

752

753

754

755

756

757

758

759

760

761

762

763

764

765

766

767

768

769

770

771

772

773

774

775

776

777

778

779

780

781

NC

OR

RE

IO 2594 1–18

B. DasGupta et al. / Bi

Some terminology The following notation will be usedor the remainder of the paper. Given a set S of vertices indirected graph G, define Eout(S) = {(u, v) ∈ E(G)|u ∈} as the set of out-bound edges of vertices in S. OPTP (I)enotes the size of an optimal solution for a problem Pith instance I. Recall that the length of a circuit c isormally defined as the number of edges in the circuit.iven a weight function w : E �→ R, the length of c with

espect to w is defined as∑

e∈c w(e).

. Theoretical results

Our theoretical results are summarized as follows.

heorem 1.

a) For some constant ε > 0, it is not possible to approx-imate in polynomial time the ULP and the DLPproblems to within an approximation ratio of 1 − ε

and 1 + ε, respectively, unless P = NP .b) For ULP, we provide a polynomial time α-

approximation algorithm where α ≈ 0.87856 is theapproximation factor for the MAX-CUT problemobtained in Goemans and Williamson (1995) viasemidefinite programming.

c) For DLP, if dmaxin denotes the maximum in-degree of

any vertex in the graph, then we give a polynomial-time approximation algorithm with an approxima-tion ratio of at most dmax

in · O(log |V |).ur computational results are illustrated in Section 6 by

n implementation of the algorithms applied to a 13-ode Drosophila segmentation network, as well as to a00+ node recently published network of the Epidermalrowth Factor Receptor pathway.

emark 1. It should be noted that the complexity ofLP becomes tractable if the network is biased signifi-

antly towards excitatory connections. Obviously, if allhe edges of the given graph G = (V, E) are labeled 0,hen it is possible to label the vertices such that all thedges are consistent. Moreover, given any graph G, its easy to check in O((|V | + |E|)3) time if an optimalolution contains all the edges as consistent by solvingset of linear equations via Gaussian elimination. Now,

uppose that at most L of the edges of G are labeled. Then, obviously at most L inconsistent edges existn any optimal solution. Thus a straightforward way toolve the problem is to consider all possible subsets ofdges in which at most L edges are dropped and check-

Ung, for each such subset, if there is an optimal solutionhat contains all the edges as consistent. The total timeaken is O(|V |2L. · (|V | + |E|)3), which is a polynomialn |V | + |E| if L is a constant.

Please cite this article as: Bhaskar DasGupta et al., Algorithmic and cinto monotone subsystems, BioSystems (2006), doi:10.1016/j.bios

ED

PR

OO

F

s xxx (2006) xxx–xxx 9

5. Proof of Theorem 1

This section provides the proof of Theorem 1, brokenup into a series of technical parts.

5.1. Proof of Theorem 1(a)

Based on the discussion in Section 3.1, it sufficesto show that both these problems are MAX-SNP-hard.ULP is MAX-SNP-hard since its special case, the MAX-CUT problem, is MAX-SNP-hard. To prove MAX-SNP-hardness of DLP, we need the definitions of the followingtwo problems.

Problem 3 (Node deletion problem with bipartite prop-erty (NDBP)). An instance of this problem is an undi-rected graph G = (V, E). A valid solution is a vertexset S ⊆ V , such that G(V − S) is a bipartite graph. Theobjective is to find a valid solution minimizing |S|.Problem 4 (Variance of node deletion problem(VNDP)). An instance of this problem is (G, h) whereG = (V, E) is a directed graph and h : E → {0, 1}. Avalid solutions is a vertex set S ⊆ V with the followingproperty: if GS = (VS, ES) is the graph with VS = V

and ES = E − Eout(S), then GS is free of odd lengthcircuit with respect to weight function h. The objectiveis to find a valid solution minimizing |S|.First, we note that DLP is equivalent to VNDP. If oneidentifies the solution set S in UNDP with the solutionset g(E − F ) in DLP, then the set of consistent edges Fin DLP corresponds to the ES in UNDP since every edge(u, v) ∈ F satisfying h(u, v) ≡ (f (u) + f (v)) (mod 2) isequivalent to stating that GS is free of odd length circuitwith respect to weight function h.

Thus, to prove the MAX-SNP-hardness of DLP itsuffices to prove that of VNDP. NDBP is known to beMAX-SNP-hard (Lund and Yannakakis, 1993). We pro-vide a L-reduction from NDBP to VNDP. For an instanceof VNDP with graph G = (V, E), construct an instanceof DLP with instance (G′, h) as follows (note that G′ isa digraph):

V ′ = V (G′) = V ∪ {Au,v, Bu,v|{u, v} ∈ E},E′ = E(G′)

= {(u, Au,v), (Au,v, Bu,v), (v, Bu,v)|{u, v} ∈ E},and h(e) = 1 for all e ∈ E′ Now, the following

BIO 2594 1–18

holds: 782

(1) If S is a solution to NDBP, it is also a solution 783

to the generated instance of UNDP. The reason 784

omplexity results for decompositions of biological networksystems.2006.08.001

Page 10: Algorithmic and complexity results for decompositions of ...dasgupta/resume/publ/papers/...UNCORRECTED PROOF BIO 2594 1–18 BIO25941–18 BioSystems xxx (2006) xxx–xxx Algorithmic

T

oSystems xxx (2006) xxx–xxx

785

786

787

788

789

790

791

792

793

794

795

796

797

798

799

800

801

802

803

804

805

806

807

808

809

810

811

812

813

814

815

816

817

818

Select a uniformly random vector r in the|V |-dimensional unit sphere and set{

0 if r · xv ≥ 0

819

820

821

822

823

824

825

826

827

828

829

830

831

832

833

834

835

836

837

838

839

840

841

842

843

844

845

846

847

848

849

CO

RR

EC

BIO 2594 1–18

10 B. DasGupta et al. / Bi

is as follows. Notice that every odd length (resp.,even length) circuit C in G corresponds to an oddlength (resp., even length) circuit C′ in G′ withrespect to the weight function h. Since G(V − S)is a bipartite graph, it is free of odd length circuits.So for each odd length cycle C of G, there existsu ∈ S such that the deletion of all out-bound edgesof u in G′ breaks its corresponding odd length cycleC′.

(2) If S′ is a solution to UNDP, then we can constructa solution S of NDBP in the following manner: foreach x ∈ S′:

if x = Au,v, add u to T ; if x = Bu,v, add v to T ;

if x = u or x = v, add xto T.

It is now easy to see that since the graph GS′ is free ofodd length circuit with respect to h, G(V − S) has noodd length circuit either.

Hence, we have OPTUNDP(G′, h) ≤ OPTNDBP(G).Moreover, given a solution S′ of UNDP, we are ableto generate a solution S of NDBP such that

||S| − OPTNDBP(G)| ≤ ||S′| − OPTUNDP(G′, h)|.Thus, our reduction satisfies Definition 1 of a L-reduction with a = b = 1.

5.2. Proof of Theorem 1(b)

Our algorithm for ULP uses the semidefinite pro-gramming (SDP) technique used by Goemans andWilliamson in Goemans and Williamson (1995); hencewe use notations and terminologies similar to that usedin the paper (readers not very familiar with this tech-nique are also referred to the excellent explanation ofthis technique in the book by Vazirani Vazirani (2001)).For each vertex v ∈ V , we have a real vector xv ∈ R|V |with ||xv||2 = 1. Then, we can generate from ULP thefollowing vector program (where · denotes the vectorinner product):

Solve the following vector program via SDPmethods:

UNmaximize

1

2

∑h(u,v)=1

(1−xu · xv)+1

2

∑h(u,v)=0

(1+xu · xv)

subject to : for each v ∈ V : xv · xv = 1for each v ∈ V

: xv ∈ R|V |.

850

Please cite this article as: Bhaskar DasGupta et al., Algorithmic andinto monotone subsystems, BioSystems (2006), doi:10.1016/j.bios

ED

PR

OO

F

f (v) =1 otherwise

This proof of the claimed approximation performanceof the above vector program is obtained by adapting theproof in Section 26.5 of Vazirani (2001) for the MAX-2SAT problem to deal with fact that, in our problem,aij = bij = 1/2 as opposed to a different set of values inVazirani (2001). Since there are some subtleties in adapt-ing that proof for readers unfamiliar with this approach,we provide a sketch of the proof in Appendix A. The pro-cedure can be derandomized via methods of conditionalprobabilities (e.g., see Mahajan and Ramesh (1995)).

5.3. Proof of Theorem 1(c)

For an instance of (G, h) of DLP, construct instance(G′ = (V ′, E′), h′) as follows:

V ′ = V ∪ {Cu,v|(u, v) ∈ E & h(u, v) = 0},E′ = {e|e ∈ E & h(e) = 1} ∪ {(u, Cu,v),

× (Cu,v, v)|(u, v) ∈ E & h(u, v) = 0},and

h′(e) = 1for alle ∈ E′.

Note that every odd (resp., even) length circuit in G withrespect to weight function h corresponds to an odd (resp.,even) length circuit in G′ with respect to weight functionh′, and vice versa. Let F is a set of consistent edges in(G, h) with a vertex labeling function f. Now, observethe following:

(1) F ′ is a set of consistent edges in (G′, h′) with avertex labeling function f ′ with f ′(x) = f (x) forx ∈ V ′ ∩ V and f ′(Cu,v) = f (u) = f (v) for an edge(u, v) ∈ F with h(u, v) = 0; thus, an edge (u, v) inF correspond to an edge (u, v) in F ′ if h(u, v) = 1and correspond to a pair of edges (u, Cu,v), (Cu,v, v)in F ′ if h(u, v) = 0.

(2) If (u, v) ∈ E − F is an inconsistent edge in (G, h),′

BIO 2594 1–18

then the edge (Cu,v, v) in G can always be made 851

consistent by choosing f ′(Cu,v) = f (v). 852

Thus, if F ′′ is the set of consistent edges obtained from F 853

following rules (1) and (2) above, then |g(E′ − F ′′)| = 854

complexity results for decompositions of biological networksystems.2006.08.001

Page 11: Algorithmic and complexity results for decompositions of ...dasgupta/resume/publ/papers/...UNCORRECTED PROOF BIO 2594 1–18 BIO25941–18 BioSystems xxx (2006) xxx–xxx Algorithmic

UN

CO

RR

EC

T

BIO 2594 1–18

B. DasGupta et al. / BioSystem

|g(E − F )| and thus OPTDLP(G′, h′) = OPTDLP(G, h).855

Consider the NDBP problem on G′. Any solution to DLP856

on (G′, h′) with vertex labeling function f ′ and set of857

consistent edges F ′ cannot contain an odd cycle of con-858

sistent edges and thus provides a solution to NDBP on859

G′ of size |g(E′ − F ′)|. Thus,860

OPTNDBP(G′) ≤ OPTDLP(G′, h′) = OPTDLP(G, h).861

OPTNDBP(G′) can be approximated in polynomial time862

to within an approximation ratio of O(log |V ′|) (Lund863

and Yannakakis, 1993), i.e., we can find a solution864

SNDBP(G′) in polynomial time such that865

|SNDBP(G′)| ≤ O(log |V ′|) · OPTNDBP(G′)866

≤ O(log |V |) · OPTDLP(G, h).867

Now,868

SDLP(G, h) = SNDBP(G′)869

× ∪ {u | ∃v ∈ SNDBP(G′), (u, v) ∈ E},870

is obviously a solution to DLP on (G, h). Recall that871

dmaxin denotes the maximum in-degree of any vertex in872

G. Thus,873

|SDLP(G, h)| ≤ dmaxin · |SNDBP(G′)|874

≤ dmaxin · O(log |V |) · OPTDLP(G, h).875

876

6. Examples of applications of the ULP877

algorithm878

We have implemented the SDP-based algorithm for879

calculating approximate solutions of the undirected880

labeling problem using Matlab, and we illustrate this881

Fig. 5. The network associated to the Drosophila segment polarity, as proposethree edges that have been crossed have been chosen in order to let the remain

882

883

884

885

886

887

888

889

890

891

892

893

894

895

896

897

898

899

900

901

902

903

904

905

906

907

908

909

Please cite this article as: Bhaskar DasGupta et al., Algorithmic and cinto monotone subsystems, BioSystems (2006), doi:10.1016/j.bios

ED

PR

OO

F

s xxx (2006) xxx–xxx 11

algorithm with two applications to biological systems.The first application concerns the relatively small-scale13-variable digraph of a model of the Drosophila seg-ment polarity network. A second application involves adigraph with 300+ variables associated to the humanEpidermal Growth Factor Receptor (EGFR) signalingnetwork. This model was published recently and builtusing information from 242 published papers. Finally,we provide an example involving a yeast gene regula-tory network.

6.1. Drosophila segment polarity

An important part of the development of the earlyDrosophila (fruit fly) embryo is the differentiation ofcells into several stripes (or segments), each of whicheventually gives rise to an identifiable part of the bodysuch as the head, the wings, the abdomen, etc. Each seg-ment then differentiates into a posterior and an anteriorpart, in which case the segment is said to be polarized.(This differentiation process continues up to the pointwhen all identifiable tissues of the fruit fly have devel-oped.) Differentiation at this level starts with differingconcentrations of certain key proteins in the cells; theseproteins form striped patterns by reacting with each otherand by diffusion through the cell membranes.

A model for the network that is responsible for seg-ment polarity (von Dassow et al., 2000) is illustratedin Fig. 5. As explained above, this model is best stud-ied when multiple cells are present interacting with each

BIO 2594 1–18

d in von Dassow et al. (2000), Courtesy of N. Ingolia and PLoS. Theing edges form an orthant monotone system.

other. But it is interesting at the one-cell level in its own 910

right—and difficult enough to study that analytic tools 911

seem mostly unavailable. The arrows with a blunt end 912

are interpreted as having a negative sign in our notation.

omplexity results for decompositions of biological networksystems.2006.08.001

Page 12: Algorithmic and complexity results for decompositions of ...dasgupta/resume/publ/papers/...UNCORRECTED PROOF BIO 2594 1–18 BIO25941–18 BioSystems xxx (2006) xxx–xxx Algorithmic

TED

PR

OO

F

oSystems xxx (2006) xxx–xxx

913

914

915

916

917

918

919

920

921

922

923

924

925

926

927

928

929

930

931

932

933

934

935

936

937

938

939

940

941

942

943

944

945

946

947

948

949

950

951

952

953

954

955

956

957

958

959

960

ics of the system. This is especially the case given that 961

the open loop digraph has almost no closed oriented 962

paths (except for WG ↔ wg), which is evidence that 963

the dynamics of the control system under constant inputs 964

may be especially simple, e.g. such that all solutions con- 965

verge towards a unique equilibrium. 966

6.1.1. Multiple copies 967

It was mentioned above that the purpose of this 968

network is to create striped patterns of protein con- 969

centrations along multiple cells. In this sense, it is 970

most meaningful to consider a coupled collection 971

of networks as it is given originally in Figs. 6 and 5. 972

Consider a row of k cells, each of which has independent 973

concentration variables for each of the compounds, and 974

let the cell-to-cell interactions be as in Fig. 5 with cyclic 975

boundary conditions (that is, the kth cell is coupled 976

with the first in the natural way). We show that the 977

results can be extended in a very similar manner as 978

before. 979

Given a partition fV of the one-cell network consid- 980

ered above, let fV be the partition of the k-cell network 981

defined by fV (eni) := fV (en) for every i, etc. Thus fV 982

consists of k copies of the partition fV in a natural way. 983

Lemma 6. Let fV be a partition of the nodes of the 1- 984

cell network with n consistent edges. Then with respect 985

Fig. 6. A diagram of the Drosophila embryo during early development.Each hexagon represents a cell containing a copy of the network in Fig.

N

CO

RR

EC

BIO 2594 1–18

12 B. DasGupta et al. / Bi

Furthermore, the concentrations of the membrane-boundand inter-cell traveling compounds PTC, PH, HH andWG (membrane) on all cells have been identified inthe one-cell model (so that, say, HH→ PH is now inthe digraph). Finally, PTC acts on the reaction CI→CN itself by promoting it without being itself affected,which in our notation means PTC→+ CN and PTC→−CI.

The implementation. The Matlab implementation ofthe algorithm on this digraph with 13 nodes and 20 edgesproduced several partitions with as many as 17 consistentedges. One of these possible partitions simply consistsof placing the three nodes ci, CI and CN in one set andall other nodes in the other set, whereby the only incon-sistent edges are CL→+ wg, CL→+ ptc, and PTC→+CN. But note that it is desirable for the resulting openloop system to have as simple remaining loops as possi-ble after eliminating all inconsistent edges. In this case,the remaining directed loops

EN−→ ci

+→ CI+→ CN

−→ en+→ EN

EN−→ ci

+→ CI+→ CN

−→ wg+→

WG+→ WG (membrane)

+→ en+→ EN

can still cause difficulties.A second partition which generated 17 consistent

edges is that in which EN, hh, CN, and the membranecompounds PTC, PH, HH are on one set, and the remain-ing compounds on the other. The edges cut are ptc→+PTC, CI→+ CN and en→+ EN, each of which elim-inates one or several positive loops. By writing theremaining consistent digraph in the form of a cascade, itis easy to see that the only loop whatsoever remaining iswg ↔ WG; this makes the analysis proposed in Encisoand Sontag (2006) easier.

In this relatively low dimensional case we can provethat in fact OPT = 17, as the results below will show.

Lemma 5. Any partition of the nodes in the digraph inFig. 5 generates at most 17 consistent edges.

Proof. From Lemma 3, a simple way to prove this state-ment is by showing that there are three disjoint cycleswith odd weighted length in the network associated toFig. 5 (disjoint in the sense that no edge is part of morethan one of the cycles). Such three disjoint cycles existin this case, and they are CI-CN-wg, CI-ptc-PTC, CN-en-EN-hh-HH-PH-PTC. �

U

BIO 2594 1–18

It is surprising that a realistic biological system with asmany as 13 variables and 20 edges can be transformedinto a monotone system after the deletion of only 3 nodes.It is conceivable that this restricts the possible dynam-

6, and neighboring cells interact to form a collective behavior. In thisexample, an initial striped pattern of the genes en and wg induces theproduction of the gene hh, but only in those cells that are producing en.This will further strengthen the pattern of stripes and help differentiatethe various tissues. Courtesy of N. Ingolia and PLoS (Ingolia, 2004).

Please cite this article as: Bhaskar DasGupta et al., Algorithmic and complexity results for decompositions of biological networksinto monotone subsystems, BioSystems (2006), doi:10.1016/j.biosystems.2006.08.001

Page 13: Algorithmic and complexity results for decompositions of ...dasgupta/resume/publ/papers/...UNCORRECTED PROOF BIO 2594 1–18 BIO25941–18 BioSystems xxx (2006) xxx–xxx Algorithmic

UN

CO

RR

EC

T

BIO 2594 1–18

B. DasGupta et al. / BioSystem

to the partition fV , there are exactly kn consistent edges986

for the k-cell coupled model.987

Proof. Consider the network consisting of k isolated988

copies of the network, that is, k groups of nodes each of989

which is connected exactly as in the one-cell case. Under990

the partition fV , this network has exactly kn consistent991

edges. To arrive to the coupled network, it is sufficient to992

replace all edges of the form (HHi, PHi) by (HHi+1, PHi)993

and (WGi, eni) by (WGi+1, eni), i = 1, . . . , k (where we994

identify k + 1 with 1). Since by definition fV (HHi+1) =995

fV (HHi) and fV (WGi+1) = fV (WGi), the consistency996

of these edges does not change, and the number of con-997

sistent edges therefore remains constant. �998

In particular, OPT≥ 17k for the coupled system. The999

following result will establish an upper bound for OPT.1000

Lemma 7. Any partition of the nodes in the digraph in1001

the k-cell coupled network generates at most 17k con-1002

sistent edges.1003

Proof. Consider the signed graph in Fig. 7, which is a1004

sub-digraph of the network associated to Fig. 5. Since1005

the inter-cell edges (WGmem,en) and (HH,PH) are not1006

in this graph, it follows that there are k identical copies1007

of it in the k-cell model. If it is shown that at least three1008

edges need to be cut in each of these k sub-digraphs, the1009

result follows immediately.1010

Consider the negative cycle ci-CI-wg-CN-en-EN,1011

which must contain at least one inconsistent edge for1012

Fig. 7. A sub-digraph of the network in Fig. 5, using the notationdefined in the previous sections. Note that this sub-digraph does notinclude any of the two edges (WGmem,en) and (HH,PH), which con-nect the networks of different cells in Fig. 5; this will be important inthe proof of Lemma 7.

1013

1014

1015

1016

1017

1018

1019

1020

1021

1022

1023

1024

1025

1026

1027

1028

1029

1030

1031

1032

1033

1034

1035

1036

1037

1038

1039

1040

1041

1042

1043

1044

1045

1046

1047

1048

1049

1050

1051

1052

1053

1054

1055

1056

1057

Please cite this article as: Bhaskar DasGupta et al., Algorithmic and cinto monotone subsystems, BioSystems (2006), doi:10.1016/j.bios

ED

PR

OO

F

s xxx (2006) xxx–xxx 13

any given partition. The remaining edges of the subgraphform a tetrahedron with four negative parity triangles,which cannot all be cut by eliminating any single edge.If follows that no two edges can eliminate all negativeparity cycles in this signed graph, and that therefore20k − 3k = 17k is an upper bound for the number ofconsistent edges in the k-cell network.

Corollary 1. For the k-cell linearly coupled networkdescribed in Fig. 5, it holds OPT = 17k.

Proof. Follows from the previous two results. �

6.2. EGFR signaling

The protein called epidermal growth factor is fre-quently stored in epithelial tissues such as skin, and it isreleased when rapid cell division is needed (for instance,it is mechanically triggered after an injury). Its functionis to bind to a receptor on the membrane of the cells, aptlycalled the epidermal growth factor receptor. The EGFR,on the inner side of the membrane, has the appearance ofa scaffold with dozens of docks to bind with numerousagents, and it starts a reaction of vast proportions at thecell level that ultimately induces cell division.

In their May 2005 paper (Oda et al., 2005), Odaet al. integrate the information that has become avail-able about this process from multiple sources, and theydefine a network with 330 known molecules under211 chemical reactions. The network itself is availablefrom supplementary material in SBML format (SystemsBiology Markup Language, http://www.sbml.org), andwill most likely be subject to continuous updates. Theimplementation. Each reaction in the network classifiesthe molecules as reactants, products, and/or modifiers(enzymes). This information was imported into Matlabusing the Systems Biology Toolbox. The digraph G thatis used for this analysis has many more edges than thedigraph considered in the digraph displayed in Oda et al.(2005). The reason for this is as follows: if molecules Aand B are both reactants in the same reaction, then thepresence of A will have an indirect inhibiting effect on theconcentration of B, since it will accelerate the reactionwhich consumes B (assuming B is not also a product).Therefore a negative edge must also appear from A to B,and vice versa. Similarly, modifiers have an inhibitingeffect on reactants.

We thus define G by letting sign(i, j) = 1 if thereexists a reaction in which j is a product and i is either

BIO 2594 1–18

a reactant or a modifier. We let sign(i, j) = −1 if there 1058

exists a reaction in which j is a reactant, and i is also 1059

either a reactant or a modifier. Similarly sign(i, j) = 0 1060

if the nodes i, j are not simultaneously involved in any 1061

omplexity results for decompositions of biological networksystems.2006.08.001

Page 14: Algorithmic and complexity results for decompositions of ...dasgupta/resume/publ/papers/...UNCORRECTED PROOF BIO 2594 1–18 BIO25941–18 BioSystems xxx (2006) xxx–xxx Algorithmic

T

oSystem

1062

1063

1064

1065

1066

1067

1068

1069

1070

1071

1072

1073

1074

1075

1076

1077

1078

1079

1080

1081

1082

1083

1084

1085

1086

1087

1088

1089

1090

1091

1092

1093

1094

1095

1096

1097

1098

1099

1100

1101

1102

1103

1104

1105

1106

1107

1108

1109

1110

1111

1112

1113

1114

1115

1116

1117

1118

1119

1120

1121

1122

1123

1124

1125

1126

1127

1128

1129

1130

1131

1132

1133

1134

1135

1136

1137

1138

1139

1140

1141

1142

1143

1144

1145

1146

1147

1148

1149

1150

1151

1152

1153

1154

1155

1156

1157

1158

1159

1160

1161

NC

OR

RE

C

BIO 2594 1–18

14 B. DasGupta et al. / Bi

given reaction, and sign(i, j) is undefined (NaN) if thefirst two conditions above are both satisfied.

In a few of the reactions of this network there is amodifier or a reactant involved which has an inhibitoryeffect in the reaction. The effect of this compound onthe remaining participants of the reaction is the oppositefrom that described above. Determining which com-pounds were inhibitors in the reaction was difficult giventhe nature of this dataset. Therefore the digraph was cor-rected by hand in this implementation by looking at theannotations given for each reaction.

An undefined edge can be thought of as an edge that isboth positive and negative, and it can be dealt with, givenan arbitrary partition, by deleting exactly one of the twosigned edges so that the remaining edge is consistent.Thus, in practice, one can consider undefined edges asedges with sign 0, and simply add the number of unde-fined edges to the number of inconsistent edges in theend of each procedure, in order to form the total numberof inputs. This is the approach followed here; there areexactly seven such entries in the digraph G.

The results. After running the algorithm several hun-dred times for this problem, and choosing that partitionwhich produced the highest number of consistent edges,the induced consistent set contained 636 out of 855 edges(ignoring the edges on the diagonal and the 7 undefinededges). See supplementary material for the relevant Mat-lab functions that carry out this algorithm. A procedureanalogous to that carried out for system (5) allows todecompose the system as the feedback loop of a con-trolled monotone system using 855 − 636 = 219 inputs.Since the induced consistent set is maximal by definition,Proposition 2 guarantees that the function h is a negativefeedback.

Contrary to the previous application, many of thereactions involve several reactants and products in a sin-gle reaction. This induces a denser amount of negativeand positive edges: even though there are 211 reactions,there are 855 (directed) edges in the 330 × 330 graph G.It is very likely that this substantially decreases OPT forthis system.

The approximation ratio of the SDP algorithm is guar-anteed to be at least 0.87 for some r, which gives theestimate OPT≤≈ 636/0.87 ≈ 731 (valid to the extentthat r has sampled the right areas of the 330-dimensionalsphere, but reasonably accurate in practice).

One procedure that can be carried out to lower thenumber of inputs is a hybrid algorithm involving out-

Uhubs, that is, nodes with an abnormally high out-degree.Recall from the description of the DLP algorithm that allthe out-edges of a node xi can be potentially cut at theexpense of only one input u, by replacing all the appear-

Please cite this article as: Bhaskar DasGupta et al., Algorithmic andinto monotone subsystems, BioSystems (2006), doi:10.1016/j.bios

ED

PR

OO

F

s xxx (2006) xxx–xxx

ances of xi in fj(x), j �= i, by u. We considered the knodes with the highest out-degrees, and eliminated allthe out-edges associated to these hubs from the reactiondigraph to form the graph G1. Then we run the ULPalgorithm on G1 to find a partition fV of the nodes anda set of m edges that can be cut to eliminate all remain-ing negative closed chains. Finally, we put back on thedigraph those edges that were taken in the first step, andwhich are consistent with respect to the partition fV . Theresult is a decomposition of the system as the negativefeedback loop of a controlled monotone system, usingat most k + m edges.

An implementation of this algorithm with k = 60yielded a total maximum number of inputs k + m = 136.This is a significant improvement over the 226 inputsin the original algorithm. Clearly, it would be worth-while to investigate further the problem of designingefficient algorithms for the DLP problem to generateimproved hybrid algorithmic approaches. The approx-imation ratios in Theorem 1(c) are not very satisfactorysince dmax

in and log |V | could be large factors; hencefuture research work may be carried out in designingbetter approximation algorithms.

We conclude with another, more tentative way to dras-tically reduce the number of inputs necessary to writethis system as the negative closed loop of a controlledmonotone system. The idea is to make suitable changesof variables in the original system using the mass conser-vation laws. Such changes of variables are discussed inmany places, for example in Volpert et al. (2000), Angeliand Sontag (2003). In terms of the associated digraph,the result of the change of variables is often the elimina-tion of one of the closed chains. The simplest target fora suitable change of variables is a set of three nodes thatform part of the same chemical reaction, for instance tworeactants and one product, or one reactant, one productand one modifier. It is easy to see that such nodes areconnected in the associated digraph by an odd lengthtriangle of three edges.

In order to estimate the number of inputs that canpotentially be eliminated by suitable changes of vari-ables, we counted pairwise disjoint, odd length trianglesin the digraph of the EGFR network. Using a greedy algo-rithm to find and tag disjoint negative feedback triangles,we found a maximal number of them in the subgraphassociated to each of the 211 chemical reactions. Specialcare was taken so that any two triangles from differentreactions were themselves disjoint. After carrying out

BIO 2594 1–18

this procedure we found 196 such triangles in the EGFR 1162

network. This is a surprisingly high number, considering 1163

that each of these triangles must have been opened in the 1164

ULP algorithm implementation above and that therefore 1165

complexity results for decompositions of biological networksystems.2006.08.001

Page 15: Algorithmic and complexity results for decompositions of ...dasgupta/resume/publ/papers/...UNCORRECTED PROOF BIO 2594 1–18 BIO25941–18 BioSystems xxx (2006) xxx–xxx Algorithmic

CT

B

oSystem

e1166

t1167

i1168

m1169

v1170

s1171

61172

1173

S1174

M1175

n1176

n1177

t1178

i1179

e1180

1181

a1182

f1183

o1184

1185

g1186

o1187

c1188

e1189

t1190

1191

m1192

r1193

s1194

a1195

a1196

i1197

d1198

11199

g1200

1201

l1202

t1203

t1204

e1205

t1206

h1207

a1208

n1209

i1210

d1211

o1212

d1213

11214

11215

1216

1217

1218

1219

1220

1221

1222

1223

1224

1225

• 1226

1227

• 1228

1229

• 1230

1231

1232

1233

1234

1235

1236

1237

1238

this provides a solution for the vector program with an 1239

objective value of precisely |FOPT|. Thus, it suffices if 1240

we prove our claim on the approximation ratio relative 1241

to SDPOPT. 1242

Next, note that the vector program can indeed be 1243

solved by a SDP approach. Let Y ∈ R|V |×|V | be an 1244

NC

OR

RE

IO 2594 1–18

B. DasGupta et al. / Bi

ach triangle must contain 1 of the 226 edges cut. Tohe extent to which most of these triangles can be elim-nated by suitable changes of variables, this can yield a

uch lower number of edges to cut, and it could pro-ide a way to thus stress the underlying structure of theystem.

.3. A yeast regulatory network

As a final example, we run our algorithm on the yeastaccharomyces cerevisiae gene regulatory network fromilo et al. (2002), downloaded from Anon (2006). This

etwork has 690 nodes and 1082 edges, of which 221 areegative and 861 are positive (we labeled the one “neu-ral” edge as positive; the conclusions will not changef we labeled it negative instead, or we deleted this onedge).

Our algorithm (with 200 randomizations) providesn answer of 43 inconsistent edges, for the best partitionound. In other words, it shows that deleting a mere 4%f edges makes the network consistent.

Also interesting is the following fact. The originalraph has 11 components: a large one of size 664, onef size 5, three of size 3, and six of size 2. All of theseomponents remain connected after edge deletion. Thedges deleted all belong to the largest component, andhey are incident on a total of 65 nodes in this component.

To better appreciate if this small number of deletionsight arise by chance, we also run our algorithm on

andom graphs having 690 nodes and 1082 edges (cho-en uniformly), of which 221 edges (chosen uniformly)re negative. We found that, for such random graphs,bout 12.6% (136.6 ± 5) of edges have to be removedn order to achieve consistency. Thus, the number ofeletions needed in the biological network is roughly5 standard deviations away from the mean for randomraphs.

It would appear that both the topology (i.e., the under-ying graph) and the actual sign assignments contributeo this near-consistency of the yeast network. To jus-ify this remark, we performed the following numericalxperiment. We randomly changed the signs of 50 posi-ive and 50 negative edges, thus obtaining a network thatas the same number of positive and negative edges,nd the same underlying graph, as the original yeastetwork, but with 100 edges, picked randomly, hav-ng different signs. Now, one needs 8.2% (88.3 ± 7.1)eletions, an amount in-between that obtained for the

Uriginal yeast network and the one obtained for ran-om graphs. Changing more signs, 100 positives and00 negatives, leads to a less consistent network, with15.4 ± 4.0 required deletions, or about 10.7% of the

Please cite this article as: Bhaskar DasGupta et al., Algorithmic and cinto monotone subsystems, BioSystems (2006), doi:10.1016/j.bios

ED

PR

OO

F

s xxx (2006) xxx–xxx 15

original edges, although still not as many as for a randomnetwork.

Appendix A. More details on SDP algorithm

In this appendix, we provide details regarding theproof of the SDP algorithm for Theorem 1(b) describedin Section 5.2. The proof method is similar to that usedin better-known problems. For simplicity, we do notdescribe the derandomization methods and provide aproof for the expected approximation ratio only. Definethe following notations for convenience:

The vertex set V of the graph for ULP is simply{1, 2, . . . , |V |};fOPT is an optimal vertex labeling for ULP with FOPTbeing the set of consistent edges;SDPOPT is the maximum value of the objective valueof the vector program

maximize1

2

∑h(u,v)=1

(1 − xu · xv) + 1

2

∑h(u, v)

= 0(1 + xu · xv)

subject to : for eachv ∈ V : xv · xv = 1

for eachv ∈ V : xv ∈ R|V |

It is easy to see that SDPOPT ≥ |FOPT| as follows. Forevery v ∈ V if fOPT(v) = 0 then set

xv = (1, 0, 0, . . . , 0︸ ︷︷ ︸|V |−1|

),

whereas if fOPT(v) = 1 then set

xv = (−1, 0, 0, . . . , 0︸ ︷︷ ︸|V |−1|

);

BIO 2594 1–18

unknown real matrix with yi,j denoting the (i, j)th ele- 1245

ment of Y. It is not difficult to see (via Cholesky decom- 1246

position for real symmetric matrices) that the above vec- 1247

tor program is equivalent to the followingsemidefinite

omplexity results for decompositions of biological networksystems.2006.08.001

Page 16: Algorithmic and complexity results for decompositions of ...dasgupta/resume/publ/papers/...UNCORRECTED PROOF BIO 2594 1–18 BIO25941–18 BioSystems xxx (2006) xxx–xxx Algorithmic

T

oSystem

1248

1249

1250

1251

1252

1253

1254

1255

1256

1257

1258

1259

1260

1261

1262

1263

1264

1265

1266

1267

1268

1269

1270

1271

1272

1273

1274

1275

1276

1277

1278

1279

1280

1281

1282

1283

1284

1285

1286

1287

1288

1289

1290

1291

1292

1293

1294

1295

1296

1297

1298

1299

1300

1301

1302

1303

1304

1305

1306

1307

1308

1309

1310

1311

OR

RE

C

BIO 2594 1–18

16 B. DasGupta et al. / Bi

programming problem:

maximize1

2

∑h(u,v)=1

(1 − yu,v) + 1

2

∑h(u,v)=0

(1 + yu,v)

subject to : for each v ∈ V : yv,v = 1

Y is a positive semidefinite matrix

Such a problem can be solved in polynomial timewithin an additive error of any constant ε > 0 via ellip-soid, interior-point or convex-programming methods(Alizadeh, 1995; Grotschel et al., 1988; Nesterov andNemirovskii, 1989, 1994; Vaidya, 1989).

Let θu,v denote the angle between the two vectorsxu, xv ∈ R|V | in an optimal solution of the vector pro-gram. Then, using standard trigonometric results,

SDPOPT = 1

2

∑h(u,v)=1

(1 − cos θu,v)

+ 1

2

∑h(u,v)=0

(1 + cos θu,v). (A.1)

Let W be the expected value of the number of consistentedges of ULP after we have performed the randomizedrounding step, namely the step:

select a uniformly random vector r in the |V |-dimensional unit sphere;

setf (v) ={

0 ifr · xv ≥ 0

1 otherwise

Then, via linearity of expectation, it follows that

E[W] =∑

h(u,v)=1

Pr[f (u) �= f (v)]

+∑

h(u,v)=0

Pr[f (u) = f (v)]. (A.2)

Because the vector r was chosen randomly, it is true that

Pr[f (u) �= f (v)] = θu,v

π

and Pr[f (u) = f (v)] = 1 − θu,v

π. (A.3)

UN

CThus,

E[W] =∑

h(u,v)=1

θu,v

π+

∑h(u,v)=0

(1 − θu,v

π

)

≥ ∆ ·⎡⎣1

2

∑h(u,v)=1

(1 − cos θu,v) + 1

2

∑h(u,v)=0

(1 + cos θ

Please cite this article as: Bhaskar DasGupta et al., Algorithmic andinto monotone subsystems, BioSystems (2006), doi:10.1016/j.bios

ED

PR

OO

F

s xxx (2006) xxx–xxx

where

∆ = min

{2

πmin

0≤θ≤π

θ

1 − cos θ, min

0≤θ≤π

2 − 2θπ

1 + cos θ

}can be shown to satisfy ∆ > 0.87856 using elementarycalculus.

A.1. Proof of lemma 3

Proof. Suppose that the system is monotone withrespect to ≤fV , that is,

fV (i)fV (j)fE(i, j) = 1 for all i, j, i �= j.

(by Lemma 2). Let V (G) = A ∪ B, where i ∈ A iffV (i) = 1, and i ∈ B otherwise. Note that by hypoth-esis fE(i, j) = 1 if xi, xj ∈ A or if xi, xj ∈ B. Also,fE(i, j) = −1 if xi ∈ A, xj ∈ B or vice versa. Notingthat every closed chain in G must cross an even numberof times between A and B, it follows that every closedchain has parity 1.

Conversely, let all closed chains in G have parity 1.We define a function fV as follows: consider the par-tition of V (G) induced by letting i∼j if there existsan undirected open chain joining i and j. Pick a rep-resentative ik of every equivalence class, and definefV (ik) = 1, k = 1, . . . , K. Next, given an arbitrary ver-tex i and the representative ik of its connected com-ponent, define fV (i) as the parity (+1 of −1) of anyundirected open chain joining ik with i. To see thatthis function is well defined, note that any two chainsjoining i and j can be put together into a closedchain from ik to itself, which has parity 1 by hypoth-esis. Thus the parity of both open chains must be thesame.

Let now i, j be arbitrary different vertices. If∂Fj/∂xi ≡ 0, then (2) is satisfied for i, j; otherwisethere is an edge joining i with j. By construction ofthe “potential” function fV , it holds that if fV (i) =fV (j) then fE(i, j) = 1, i.e., ∂Fj/∂xi ≥ 0, and so (2)holds as well. If fV (i) �= fV (j), then fE(i, j) = −1, i.e.∂Fj/∂xi ≤ 0. In that case (2) also holds, and the proof iscomplete. �

BIO 2594 1–18

u,v)

⎤⎦ = ∆ · SDPOPT (A.4)

complexity results for decompositions of biological networksystems.2006.08.001

Page 17: Algorithmic and complexity results for decompositions of ...dasgupta/resume/publ/papers/...UNCORRECTED PROOF BIO 2594 1–18 BIO25941–18 BioSystems xxx (2006) xxx–xxx Algorithmic

CTE

D P

RO

OF

B

oSystems xxx (2006) xxx–xxx 17

A1312

1313

c1314

j1315

R1316

A1317

1318

1319

A1320

1321

1322

1323

A1324

1325

1326

1327

A1328

1329

1330

A1331

1332

A1333

1334

A1335

1336

1337

1338

A1339

1340

1341

B1342

1343

B1344

1345

1346

1347

B1348

1349

C1350

1351

1352

C1353

1354

C1355

1356

1357

1358

D1359

1360

1361

D1362

1363

1364

D1365

1366

1367

1368

von Dassow, G., Meir, E., Munro, E.M., Odell, G.M., 2000. The seg- 1369

ment polarity network is a robust developmental module. Nature 1370

406, 188–192. 1371

DeAngelis, D.L., Post, W.M., Travis, C.C., 1986. Positive Feedback in 1372

Natural Systems. Springer-Verlag, New York. 1373

Enciso, G.A., Smith, H.L., Sontag, E.D., in press. Non-monotone 1374

systems decomposable into monotone systems with negative feed- 1375

back, J. Diff. Eq. 1376

Enciso, G., Sontag, E., 2004. On the stability of a model of testosterone 1377

dynamics. J. Math. Biol. 49, 627–634. 1378

Enciso, G., Sontag, E.D., 2005. Monotone systems under positive feed- 1379

back: multistability and a reduction theorem. Syst. Contr. Lett. 54, 1380

159–168. 1381

Enciso, G., Sontag, E., 2006. Global attractivity, I/O monotone 1382

small-gain theorems, and biological delay systems. Discr. Contin. 1383

Dynam. Syst. 14, 549–578. 1384

Gedeon, T., Sontag, E.D., 2005. Oscillation in multi-stable monotone 1385

system with slowly varying positive feedback, submitted for pub- 1386

lication (abstract in Sixth SIAM Conference on Control and its 1387

Applications, New Orleans, July). 1388

Goemans, M., Williamson, D., 1995. Improved approximation algo- 1389

rithms for maximum cut and satisfiability problems using semidef- 1390

inite programming. J. ACM 42 (6), 1115–1145. 1391

Grotschel, M., Lovasz, L., Schrijver, A., 1988. Geometric Algorithms 1392

and Combinatorial Optimization. Springer-Verlag, New York, NY. 1393

Hastad, J., Venkatesh, S., 2002. On the advantage over a random assign- 1394

ment. Proceedings of the 34th Annual ACM Symposium on Theory 1395

of Computing, 43–52. 1396

Hirsch, M., 1985. Systems of differential equations that are competitive 1397

or cooperative. II. Convergence almost everywhere. SIAM J. Math. 1398

Anal. 16, 423–439. 1399

Hirsch, M., 1983. Differential equations and convergence almost 1400

everywhere in strongly monotone flows. Contemp. Math. 17, 267– 1401

285. 1402

http://www.weizmann.ac.il/mcb/UriAlon/Papers/networkMotifs/ 1403

yeastData.mat. 1404

Ingolia, N., 2004. Topology and robustness in the drosophila segment 1405

polarity network. Publ. Libr. Sci. 2 (6), 0805–0815. 1406

Istrail, S., 2000. Statistical mechanics, three-dimensionality and np- 1407

completeness. I. Universality of intractability of the partition 1408

functions of the ising model across non-planar lattices. Proceed- 1409

ings of the 32nd ACM Symposium on the Theory of Computing 1410

(STOC00), Portland, Oregon, May 21–23, 2000. ACM Press 87– 1411

96. 1412

Klamt, S., 2006. Generalized concept of minimal cut sets in biochem- 1413

ical networks. Biosystems 83, 233–247. 1414

Lewis, J., Slack, J.M., Wolpert, L., 1977. Thresholds in development. 1415

J. Theor. Biol. 65, 579--590. 1416

Lund, C., Yannakakis, M., 1993. The approximation of maximum sub- 1417

graph problems. Proceedings of the International Colloquium on 1418

Automata, Languages and Programming, Lecture Notes in Com- 1419

puter Science, 700, Springer-Verlag. 1420

Mahajan, S., Ramesh, H., 1995. Derandomizing semidefinite program- 1421

ming based approximation algorithms. Proceedings of the 37th 1422

Annual IEEE symposium on Foundations of Computer Science, 1423

162–169. 1424

Ma’ayan, A., Iyengar, R., Sontag, E.D., in preparation. Sign- 1425

NC

OR

RE

IO 2594 1–18

B. DasGupta et al. / Bi

ppendix B. Supplementary data

Supplementary data associated with this articlean be found, in the online version, at 10.1016/.biosystems.2006.08.001.

eferences

lizadeh, F., 1995. Interior point methods in semidefinite program-ming with applications to combinatorial optimization. SIAM J.Optimiz. 5, 13–51.

maldi, E., Kann, V., 1996. On the approximability of some NP-hardminimization problems for linear systems, ECCC Report TR96–015, (available electronically from http://eccc.uni-trier.de/eccc-reports/1996/TR96–015/).

ngeli, D., Ferrell Jr., J.E., Sontag, E.D., 2004. Detection of multi-stability, bifurcations, and hysteresis in a large class of biologicalpositive-feedback systems. Proc. Natl. Acad. Sci. U.S.A. 101,1822–1827.

ngeli, D., De Leenheer, P., Sontag, E.D., 2004. A small-gain theoremfor almost global convergence of monotone systems. Syst. Contr.Lett. 51, 185–202.

ngeli, D., Sontag, E.D., 2003. Monotone control systems. IEEETrans. Autom. Contr. 48, 1684–1698.

ngeli, D., Sontag, E.D., 2004. Multistability in monotone I/O sys-tems. Syst. Contr. Lett. 51, 185–202.

ngeli, D., Sontag, E.D., 2004. An analysis of a circadian model usingthe small-gain approach to monotone systems. Proceedings of theIEEE Conference Decision and Control, Paradise Island, Bahamas,December 2004. IEEE Publications 575–578.

rora, S., Lund, C., Motwani, R., Sudan, M., Szegedy, M., 1998. Proofverification and hardness of approximation problems. J. ACM 45(3), 501–555.

arahona, F., 1982. On the computational complexity of Ising spinglass models. J. Phys. A. Math. Gen. 15, 3241–3253.

erman, P., Karpinski, M., 2001. Efficient amplifiers and boundeddegree optimization, ECCC Report TR01-053, July (available elec-tronically from http://eccc.uni-trier.de/eccc-reports/2001/TR01–053/).

erman, P., Schnitger, G., 1992. On the complexity of approximatingthe independent set problem. Inform. Comput. 96, 77–94.

inquin, O., Demongeot, J., 2002. Positive and negative feedback:striking a balance between necessary antagonists. J. Theor. Biol.216, 229–241.

ipra, B.A., 2000. The Ising Model Is NP-Complete, SIAM News, vol.33, Number 6, July/August.

reignou, N., Khanna, S., Sudan, M., 2001. Complexity classificationsof Boolean constraint satisfaction problems, SIGACT News, 32(4), Whole Number 121, Complexity Theory Column 34, 24–33,November.

e Leenheer, P., Angeli, D., Sontag, E.D., 2005. On predator-prey systems and small-gain theorems. J. Math. Biosci. Eng. 2,25–42.

e Leenheer, P., Malisoff, M., 2006. A small-gain theorem for mono-tone systems with multivalued input-state characteristics. IEEE

U

BIO 2594 1–18

Trans. Automat. Contr. 51, 287–292.e Simone, C., Diehl, M., Junger, M., Mutzel, P., Reinelt, G., Rinaldi,

G., 1995. Exact ground states of Ising spin glasses: new experi-mental results with a branch and cut algorithm. J. Stat. Phys. 80,487–496.

consistency loops detected in biochemical regulatory networks 1426

may provide dynamical stability. 1427

Meinhardt, H., 1978. Space-dependent cell determination under 1428

the control of morphogen gradient. J. Theor. Biol. 74, 307-- 1429

321. 1430

Please cite this article as: Bhaskar DasGupta et al., Algorithmic and complexity results for decompositions of biological networksinto monotone subsystems, BioSystems (2006), doi:10.1016/j.biosystems.2006.08.001

Page 18: Algorithmic and complexity results for decompositions of ...dasgupta/resume/publ/papers/...UNCORRECTED PROOF BIO 2594 1–18 BIO25941–18 BioSystems xxx (2006) xxx–xxx Algorithmic

oSystem

1431

1432

1433

1434

1435

1436

1437

1438

1439

1440

1441

1442

1443

1444

1445

1446

1447

1448

1449

1450

1451

1452

1453

1454

1455

1456

1457

1458

1459

1460

1461

1462

1463

1464

1465

1466

1467

1468

1469

1470

1471

1472

1473

1474

BIO 2594 1–18

18 B. DasGupta et al. / Bi

Milo, R., Shen-Orr, S., Itzkovitz, S., Kashtan, N., Alon, D.U., 2002.Network motifs: simple building blocks of complex networks. Sci-ence 298, 824–827.

Monod, J., Jacob, F., 1961. General conclusions: telenomic mecha-nisms in cellular metabolism, growth, and differentiation. ColdSpring Harbor Symp. Quant. Biol. 26, 389--401.

Murray, J.D., 2002. Mathematical Biology. I. An introduction.Springer, New York.

Nesterov, Y., Nemirovskii, A., 1989. Self-Concordant Functions andPolynomial Time Methods in Convex Programming. Central Eco-nomic and Mathematical Institute, USSR Academy of Science.

Nesterov, Y., Nemirovskii, A., 1994. Interior Point Polynomial Meth-ods in Convex Programming. Society of Industrial and AppliedMathematics, Philadelphia, PA.

Oda, K., Matsuoka, Y., Funahashi, A., Kitano, H., 2005. A comprehen-sive pathway map of epidermal growth factor receptor signaling.Mol. Syst. Biol. doi:10.1038/msb4100014.

Papadimitriou, C.H., Yannakakis, M., 1991. Optimization, approxima-tion, and complexity classes. J. Comput. Syst. Sci. 43 (3), 425–440.

Plathe, E., Mestl, T., Omholt, S.W., 1995. Feedback loops, stability

UN

CO

RR

EC

Tand multistationarity in dynamical systems. J. Biol. Syst. 3, 409--413.

Remy, E., Mosse, B., Chaouiya, C., Thieffry, D., 2003. A descriptionof dynamical graphs associated to elementary regulatory circuits.Bioinformatics 19 (Suppl. 2), ii172--ii178.

1475

Please cite this article as: Bhaskar DasGupta et al., Algorithmic andinto monotone subsystems, BioSystems (2006), doi:10.1016/j.bios

OO

F

s xxx (2006) xxx–xxx

Smith, H.L., 1995. Monotone Dynamical Systems. AMS, Providence,R.I.

Smith, H.L., 1988. Systems of ordinary differential equations whichgenerate an order-preserving flow: a survey of results. SIAM Rev.30, 87–111.

Snoussi, E.H., 1998. Necessary conditions for multistationarity andstable periodicity. J. Biol. Syst. 6, 3–9.

Sontag, E.D., 1990. Mathematical Control Theory: DeterministicFinite Dimensional Systems, 2nd ed. 1998. Springer, New York.

Sontag, E.D., 2004. Some new directions in control theory inspired bysystems biology. Syst. Biol. 1, 9–18.

Sontag, E.D., 2005. Molecular systems biology and control. Eur. J.Contr. 11, 396–435.

Sontag, E.D., in preparation. Consistency of indirect effects in biolog-ical networks.

Thomas, R., 1978. Logical analysis of systems comprising feedbackloops. J. Theor. Biol. 73, 631--656.

Vaidya, P., 1989. A new algorithm for minimizing convex functionsover convex sets. Proceedings of the 30th Annual IEEE Symposiumon Foundations of Computer Science 338–343.

Vazirani, V.V., 2001. Approximation Algorithms. Springer-Verlag,

ED

PR

BIO 2594 1–18

Berlin. 1476

Volpert, A.I., Volpert, V.A., Volpert, V.A., 2000. Traveling Wave Solu- 1477

tions of Parabolic Systems, volume 140 of Translations of Mathe- 1478

matical Monographs, AMS. 1479

complexity results for decompositions of biological networksystems.2006.08.001