Skip to main content

Analysis and verification of the HMGB1 signaling pathway

Abstract

Background

Recent studies have found that overexpression of the High-mobility group box-1 (HMGB1) protein, in conjunction with its receptors for advanced glycation end products (RAGEs) and toll-like receptors (TLRs), is associated with proliferation of various cancer types, including that of the breast and pancreatic.

Results

We have developed a rule-based model of crosstalk between the HMGB1 signaling pathway and other key cancer signaling pathways. The model has been simulated using both ordinary differential equations (ODEs) and discrete stochastic simulation. We have applied an automated verification technique, Statistical Model Checking, to validate interesting temporal properties of our model.

Conclusions

Our simulations show that, if HMGB1 is overexpressed, then the oncoproteins CyclinD/E, which regulate cell proliferation, are overexpressed, while tumor suppressor proteins that regulate cell apoptosis (programmed cell death), such as p53, are repressed. Discrete, stochastic simulations show that p53 and MDM2 oscillations continue even after 10 hours, as observed by experiments. This property is not exhibited by the deterministic ODE simulation, for the chosen parameters. Moreover, the models also predict that mutations of RAS, ARF and P21 in the context of HMGB1 signaling can influence the cancer cell's fate - apoptosis or survival - through the crosstalk of different pathways.

Background

The cell cycle is strictly regulated and controlled by a complex network of signaling pathways [1], comprised of hundreds of proteins. If some important proteins are mutated or there are defects in the signaling mechanisms, normal cell growth regulation will break down, possibly leading to the occurrence of cancer in the future. Moreover, a number of extracellular proteins can bind to their receptors and activate signaling pathways that promote the proliferation of cancer cells.

The high-mobility group box-1 (HMGB1) protein is a DNA-binding nuclear protein, released actively in response to cytokine stimulation, or passively during cell death [2], and it is present in almost all eukaryotic cells [3–6]. HMGB1 can activate a series of signaling components, including mitogen-activated protein kinases (MAPKs) and AKT, which play an important role in tumor growth and inflammation, through binding to different surface receptors, such as RAGE and TLR2/4. Several studies have shown that elevated expression of HMGB1 occurs in many tumors [7–10] and accelerates cell-cycle progression. Recent in vitro studies with pancreatic cancer cells [11] revealed that the targeted knockout or inhibition of HMGB1 and RAGE could increase apoptosis and suppress pancreatic cancer cell growth. This phenomenon has been also observed with lung cancer and other types of cancer cells [8, 12].

The HMGB1 signal transduction can influence the cell's fate by two important processes - apoptosis and cell proliferation - which are regulated respectively by the proteins p53 and CyclinE, acting in two different signaling pathways. The protein p53 is one of the most important tumor suppressor proteins: its activation can lead to cell cycle arrest, DNA repair, or apoptosis. Mutations of p53 occur at a frequency of 50% or higher in many different cancer types [13]. CyclinE is a cell cycle regulatory protein which regulates the G1-S phase transition during cell proliferation. Cancer cells often exhibit high expression levels of CyclinE and aberrant CyclinE activity [14]. Many studies have found evidence of crosstalk between the two signaling pathways involving p53 and CyclinE [15]. The crosstalk is regulated by tumor suppressor proteins, including ARF, P21 and FBXW7, which are also frequently mutated in many cancers. In this paper, we ask the following questions: How do these proteins and their mutations change the cell's fate - apoptosis or survival - when HMGB1 signal transduction is activated? Which signaling pathways are fundamental for describing HMGB1 signal transduction, and what mechanisms are responsible to explain recent results linking overexpression of HMGB1 with decrease of apoptosis (and increased cancer cell survival)?

To the best of the authors' knowledge, no computational model has been proposed to investigate the importance of HMGB1 in tumor proliferation. In this work, we construct a simple model of HMGB1 signal transduction to investigate tumorigenesis on the basis of known signaling pathway studies [16–21]. We also constructed a crosstalk network between these known pathways based on hypothetical mechanisms suggested by recent experiments. The HMGB1 pathway is not well understood at the mechanistic level, so our model can provide some insights into the study of HMGB1's roles in tumor proliferation. A series of deterministic and stochastic simulation experiments was conducted to investigate the properties of the HMGB1 pathway.

Finally, we analyze our pathway model against interesting behavorial properties by means of Model Checking techniques. Model Checking is an automated verification technique for hardware and software systems [22]. Recently, there has been growing interest in formal verification of stochastic systems, and, which has recently seen a growing number of applications to biological systems [23–25], by means of Model Checking techniques. The Methods section introduces statistical model checking, which we then apply to validate our pathway model against experimental results from the literature.

Methods

HMGB1 signaling pathway

Our HMGB1 signaling pathway model is illustrated in Fig. 1. It includes 31 molecular species (6 tumor suppressor proteins), 59 chemical reactions, and three different signaling pathways activated by HMGB1: the RAS-ERK, Rb-E2F and p53-MDM2 pathways. Since the interaction between HMGB1 and its receptors TLR and RAGE is not clear at the mechanistic level, RAGE is used to represent all the receptors in our model in order to reduce the number of unknown parameters. We now briefly discuss the three pathways and their crosstalk. We denote activation (or promotion) by →, while inhibition (or repression) is denoted by ⊣.

Figure 1
figure 1

Schematic view of HMGB1 signal transduction. Blue nodes represent tumor suppressor proteins; red nodes represent oncoproteins/lipids; brown node represents protein complex formed by oncoprotein E2F and tumor suppressor protein RB. Solid lines with arrows denote protein transcription, degradation or changes of molecular species; dashed lines with arrows denote activation processes.

The p53-MDM2 pathway is regulated by a negative feedback loop [26]: PI3K → PIP3 → AKT → MDM2 ⊣ p53 → MDM2, and a positive feedback loop: p53 → PTEN ⊣ PIP3 → AKT → MDM2 ⊣ p53. The protein PI3K is activated by the toll-like receptors (TLR2/4) within several minutes after TLR2/4 activation by HMGB1 [27]. In turn, PI3K phosphorylates the phosphatidylinositol 4,5-bisphosphate (PIP2) to phosphatidylinositol (3,4,5)-trisphosphate (PIP3), leading to phosphorylation of AKT. The unphosphorylated oncoprotein MDM2, which is one of p53's transcription targets [28], resides in the cytoplasm, and cannot enter the nucleus until it is phosphorylated by activated AKT. The phosphorylated MDM2 translocates into the nucleus to bind with p53, inhibiting p53's transcription activity and initializing p53 polyubiquitination [29], which targets it for degradation. Also, p53 can regulate the transcription of PTEN [30], a tumor suppressor protein, which can hydrolyze PIP3 to PIP2, thereby inhibiting the activation of AKT and MDM2.

The RAS-ERK pathway is the activation sequence: RAS → RAF → MEK → ERK → CyclinD. Activation of RAGE by HMGB1 leads to RAS activation, which in turn activates its effector protein RAF. Activated RAF will phosphorylate the MEK proteins (mitogen-activated protein kinase kinases (MAPKK)), leading to the phosphorylation of ERK1/2 (also called MAPKs). Activated ERK can phosphorylate some transcription factors which activate the expression of the regulatory protein CyclinD and Myc, enabling progression of the cell cycle through the G1 phase. K-RAS, a member of the RAS protein family, is found to be mutated in over 90% of pancreatic cancers [31].

The Rb-E2F pathway is composed of the interactions: CyclinD ⊣ Rb ⊣ E2F → CyclinE ⊣ Rb. The Rb-E2F pathway regulates the G1-S phase transition in the cell cycle during cell proliferation. E2F is a transcription factor that can activate the transcription of many proteins involved in DNA replication and cell-cycle progression [32]. In quiescent cells, E2F is bound by unphosphorylated Rb - a tumor suppressor protein - forming an Rb-E2F complex which inhibits E2F's transcription activity. E2F will be activated and released when its inhibitor Rb is phosphorylated by some oncoproteins (CyclinD and Myc in Fig. 1), leading to the transcription of CyclinE and Cyclin-dependent protein kinase 2 (CDK2) which promote cell-cycle progression. CyclinE, in turn, continues to inhibit the activity of Rb, leading to a positive feedback loop [33–35]. Fig. 1 shows that the activity of CyclinD-CDK4/6 (only CyclinD is shown in Fig. 1) is inhibited by the tumor suppressor protein INK4A, which is inactivated in up to 90% pancreatic cancers [36].

The crosstalk between these pathways can influence the cell's fate since the three signaling pathways in HMGB1 signal transduction are not independent. As shown in Fig. 1, the oncoprotein RAS can also activate the PI3K-AKT signaling pathway; the tumor suppressor ARF protein induced by E2F can bind to MDM2 to promote its rapid degradation and stabilize p53. Furthermore, it has been experimentally observed [13] that the p53-dependent tumor suppressor proteins P21 and FBXW7 can inhibit the activity of cyclin dependent kinases (In Fig. 1, we use P21 to represent both P21 and FBXW7's contribution). Mutations of RAS, ARF, P21 and FBXW7 have been found in many cancers [31, 36, 37]. One of our aims is to investigate how these mutations might influence the cell's fate.

In the HMGB1 model, all substrates are expressed in the number of molecules; proteins with the subscript "a" or "p" correspond respectively to active or phosphorylated forms of the proteins. For example,

  • RAGE (RAGE a ) - inactive (active) form of HMGB1's receptor

  • MDM2 (MDM2 p ) - unphosphorylated (phosphorylated) MDM2.

We denote the mRNA transcript of MDM2 by mdm 2. We assume that the total number of active and inactive forms of the RAGE, PI3K, PIP, AKT, RAS, RAF, MEK, and ERK molecules is constant. For example, AKT + AKT p = AKT tot , PIP2 + PIP3 = PIP tot . We sometimes use CD to stand for the CyclinD-CDK4/6 complex, CE for CyclinE, and RE for the Rb-E2F complex.

The p53-MDM2 and RAS-ERK pathways have been studied individually using deterministic ODE methods [16–19, 32]. We instead formulated a reaction model corresponding to the reactions illustrated in Fig. 1 in the form of rules specified in the BioNetGen language [38]. We used Hill functions to describe the rate laws governing protein synthesis, including PTEN, MDM2, CyclinD (CD), Myc, E2F and CyclinE (CE). Our choice was motivated by several studies [19, 39–41], which showed that transcription rates of these proteins are sigmoidal functions of transcription factor (TF) concentrations with positive cooperative Hill coefficients. We used mass action rules for other types of chemical reactions. Both ODEs and Gillespie's stochastic simulation algorithm (SSA) [42] are used to simulate the model with BioNetGen [38]. Stochastic simulation is important because when the number of molecules involved in the reactions is small, stochasticity and discretization effects become more prominent [43–45]. In the online Additional file 1, we list 23 ordinary differential equations which describe the deterministic HMGB1 model and all the input parameters. The BioNetGen code which implements SSA and ODE models is available at [46].

Since our understanding of many chemical reactions at the mechanistic level is not clear, a large number of parameters involved in these reactions are difficult to estimate based on existing data. We emphasize that in our HMGB1 model the values for some undetermined parameters were chosen in order to produce a qualitative agreement with previous experiments.

Model Checking

Model Checking [22, 47] is one of the leading techniques for the automated verification and analysis of hardware and software systems. Given a high-level behavior specification, a model checker verifies whether a system (or model) satisfies it. A specification might be satisfied by many different models. Thus, model checking is the process of determining whether or not a given system model satisfies (is a model of) a property describing the desired behavior of the system. Mathematically, system models take the form of state-transition diagrams, while some version of temporal logic [48] is used to describe the desired properties (specifications) of system executions. A typical property stated in temporal logic is G(grant_req→ F ack), meaning that it is always (G = globally) true that a grant request eventually (F = future) triggers an acknowledgment. One important aspect of Model Checking is that it can be performed algorithmically - user intervention is limited to providing a system model and a property to check.

The Probabilistic Model Checking problem (PMC) is to decide whether a stochastic model satisfies a temporal logic property with a probability greater than or equal to a certain threshold. To express temporal properties, we use a logic in which the temporal operators are equipped with bounds. For example, the property "CyclinD will always stay below 10 in the next fifty time units " is written as G50(CyclinD < 10). We now ask whether our stochastic system M satisfies that formula with a probability greater than or equal to a fixed threshold (say 0.9), and we write M |= Pr≥ 0.9[G50(CyclinD < 10)]. In the next section, we formally define the temporal logic used in this work, Bounded Linear Temporal Logic [23].

Bounded Linear Temporal Logic (BLTL)

Let SV be a finite set of real-valued variables, an atomic proposition AP be a boolean predicate of the form e1 ~ e2, where e1 and e2 are arithmethic expressions over variables in SV, and ~ is either ≥, ≤, <, >, or = . A BLTL property is built over atomic propositions using boolean connectives and bounded temporal operators. The syntax of the logic is the following:

ϕ : : = A P | ϕ 1 ∨ ϕ 2 | ϕ 1 ∧ ϕ 2 | ¬ ϕ 1 | ϕ 1 U t ϕ 2 .

The bounded until operator ϕ1 Ut ϕ2 requires that, within time t, ϕ2 will be true and ϕ1 will hold until then. Bounded versions of the F and G operators can be easily defined: Ft ϕ = true U t ϕ requires ϕ to hold true within time t;Gt ϕ = ¬Ft ¬ ϕ requires ϕ to hold true up to time t.

The semantics of BLTL is defined with respect to traces (or executions) of a system. In our case, a trace will be the output of a simulation of a BioNetGen stochastic model. Formally, a trace is a sequence of time-stamped state transitions of the form σ = (s0,t0), (s1,t1),..., which means that the system moved to state si+1after having sojourned for time t i in state s i . The fact that a trace σ satisfies the BLTL property ϕ is written as σ |= ϕ. We denote the trace suffix starting at step k by σk. We have the following semantics of BLTL:

  • σk⊨ AP if and only if AP holds true in state s k ;

  • σk⊨ Ï•1 ∧ Ï•2 if and only if σk⊨ Ï•1 and σk⊨ Ï•2;

  • σk⊨ Ï•1 ∨ Ï•2 if and only if σk⊨ Ï•1 or σk⊨ Ï•2;

  • σk⊨ ¬ϕ1 if and only if σk⊨ Ï•1 does not hold;

  • σk⊨ Ï•1UtÏ•2 if and only if there exists i ∈ N such that, (a)∑0 ≤ l <itk+l≤ t, (b) σk+i⊨ Ï•2 and (c) for each 0 ≤ j <i, σk+j⊨ Ï•1.

The semantics of BLTL are defined over infinite traces, but it can be shown that traces of an appropriate (finite) length are sufficient to decide BLTL properties [49].

Statistical Model Checking

We briefly explain Statistical Model Checking [50, 51], the technique we use for verifying BioNetGen models simulated by Gillespie's algorithm. Statistical Model Checking treats the Probabilistic Model Checking problem as a statistical inference problem, and solves it by randomized sampling of the traces (simulations) from the model. In particular, the PMC problem is naturally phrased as a hypothesis testing problem, i.e., deciding between two hypotheses - M ⊨ Pr≥θ[ϕ] versus M ⊨ Pr < θ [ϕ]. In other words, to determine whether a stochastic system M satisfies ϕ with a probability p ≥ θ, we test the hypothesis H0 : p ≥ θ against H1 : p < θ. Sampled traces are model checked individually to determine whether a given property ϕ holds, and the number of satisfying traces is used by a hypothesis testing procedure to decide between H0 and H1. Note that Statistical Model Checking cannot guarantee a correct answer to the PMC problem. However, the probability of giving a wrong answer can be made arbitrarily small.

We have introduced a Bayesian sequential hypothesis testing approach and applied it to the verification of rule-based models of signaling pathways and other stochastic systems [23, 49]. Sequential sampling means that the number of sampled traces is not fixed a priori, but is instead determined at "run-time ", depending on the evidence gathered by the samples seen so far. This often leads to a significantly smaller number of sampled traces.

Suppose that the stochastic system M satisfies the BLTL formula ϕ with some (unknown) probability p. The key idea behind statistical model checking [50] is that the behavior of M (with respect to property ϕ) can be modeled by a Bernoulli random variable with success parameter p. Such a random variable can be repeatedly evaluated via system simulation in the following way. Let σ be a trace of M, then the Bernoulli random variable X with (conditional) probability mass function:

f ( x | p ) = p x ( 1 − p ) 1 − x x ∈ { 0 , 1 }
(1)

denotes the outcome of σ ⊨ ϕ (i.e., model checking ϕ on σ). In other words, we have that:

X = { 1 with probability p ( σ | = Ï• ) , 0 with probability 1 − p ( σ | = ¬ Ï• ) .
(2)

Therefore, by running a system simulation (i.e., a BioNetGen stochastic simulation) and by checking Ï• on the resulting trace we can obtain a sample from random variable X. When a sample of X evaluates to 1 we call it a success, otherwise, a failure.

Recall that in hypothesis testing we decide between a null hypothesis H0 and an alternative hypothesis H1:

H 0 : p ≥ θ H 1 : p < θ .
(3)

The Bayesian approach assumes that p is given by a random variable whose distribution is called the prior distribution. The prior is usually based on our previous experiences and knowledge about the system.

Since p is a probability, we need prior distributions defined over [0,1]. In particular, Beta priors are mathematically convenient to use. They are defined by the following probability density:

∀ u ∈ [ 0 , 1 ] g ( u , α , β ) = 1 B ( α , β ) u α − 1 ( 1 − u ) β − 1
(4)

where the Beta function B(α, β) is defined as:

B ( α , β ) = ∫ 0 1 t α − 1 ( 1 − t ) β − 1 d t .
(5)

For later use, the Beta distribution function F(α;β)(u) of parameters α, β is defined as for all u ∈ [0, 1] as:

F ( α , β ) ( u ) = ∫ 0 u g ( t , α , β ) d t
(6)
= 1 B ( α , β ) ∫ 0 u t α − 1 ( 1 − t ) β − 1 d t .
(7)

Let d = (x1,..., x n ) denote n samples of the Bernoulli random variable X defined by (2). Let H0 and H1 be the hypotheses in (3), and suppose that the prior probabilities P (H0) and P (H1) are strictly positive and satisfy P (H0) + P (H1) = 1. By Bayes's theorem, the posterior probabilities of H0 and H1, with respect to data d, are:

P ( H i | d ) = P ( d | H i ) P ( H i ) P ( d ) ( i = 0 , 1 )

for every d with P (d) > 0. In our case, P (d) is always non-zero (there are no impossible finite sequences of data). The hypothesis test method is based on the Bayes Factor, that is, the likelihood ratio of the sampled data with respect to the two hypotheses. The Bayes Factor ℬ of sample d and hypotheses H0 and H1 is

ℬ = P ( d | H 0 ) P ( d | H 1 )

and by Bayes' theorem, we have that:

ℬ = P ( H 0 | d ) P ( H 1 | d ) · P ( H 1 ) P ( H 0 ) .
(8)

Therefore, B can be interpreted as a measure of evidence (given by the data d) in favor of H0. Now, fix a threshold T > 1. The algorithm iteratively draws independent and identically distributed (iid) sample traces in the form of BioNetGen stochastic simulations, and checks whether they satisfy ϕ (Note that BioNetGen ensures by construction that each simulation, or trace, is actually iid.) After each trace, the algorithm computes the Bayes Factor B to check if it has obtained conclusive evidence. The algorithm accepts H0 if B >T, and rejects H0 (accepting H1) if ℬ < 1 T . Otherwise ( 1 T ≤ ℬ ≤ T ) , it continues drawing iid samples. The statistical Model Checking algorithm is shown in Figure 2.

Figure 2
figure 2

Statistical Model Checking Algorithm. The algorithm for Statistical Model Checking is based on Bayesian Hypothesis Testing.

The following Proposition shows that, in our special case of Bernoulli samples, the computation of the Bayes Factor is straightforward.

Proposition 1. [49]The Bayes Factor of H0 : p ≥ θ vs. H 1 : p <θ with Bernoulli samples (x1,..., x n ) and Beta prior of parameters α, β is:

ℬ n = P ( H 1 ) P ( H 0 ) · ( 1 F ( x + α , n − x + β ) ( θ ) − 1 )

where x = ∑ i = 1 n x i is the number of successes in(x1,..., x n ) and F(s,t)(·) is the Beta distribution function of parameters s, t.

The Beta distribution function can be efficiently computed by standard software packages. Thus, no numerical integration is required for the evaluation of the Bayes Factor.

Finally, we must show that the error probability of our decision procedure, i.e., the probability that we reject (accept) the null hypothesis although it is true (false), can be bounded.

Theorem. [49]The error probability for the sequential Bayesian hypothesis testing algorithm is bounded above by 1 T where T is the Bayes Factor threshold given as input.

Results and discussion

We first conducted a series of deterministic and stochastic simulation experiments to study the properties of our HMGB1 signaling pathway model. Then, we applied the statistical model checking technique to validate some important temporal properties of our HMGB1 model.

Simulation results

We carried out a baseline simulation for four important proteins - p53, MDM2 p , CyclinD/E - using ODE and stochastic simulation. We set the initial value for the number of HMGB1 molecules to be 103; Table 1 lists all proteins with nonzero initial values; the unlisted proteins are set to 0 initially.

Table 1 Initial values for the model

The baseline stochastic simulations in Fig. 3A demonstrate that the expression levels of p53 and MDM2 p oscillate even after 10 hours, when the cell enters the S phase (recall that cells usually remain in phase G1 for about 10 hours before moving to the S phase). However, oscillations are strongly damped in the ODE simulations (Fig. 3C) when the cell proceeds to the S phase, approximately after 10 hours. The stochastic simulation model is thus more consistent with the experimental results of Geva-Zatorsky et al. [52]. In that experiment the authors measured the dynamics of p53 and MDM2 p in human breast cancer cells damaged by γ radiation. It was observed that the oscillations of p53 and MDM2 p expression levels can last more than 72 hours after irradiation.

Figure 3
figure 3

HMGB1 baseline simulation. Number of p53, MDM2 p (A, C), CyclinD/E (B, D) molecules versus time for baseline simulations with SSA (A-B) and ODE (C-D) models.

Fig. 3B and 3D show that the CyclinE protein, which regulates the G1-S phase transition in the cell cycle, reaches its maximum at about 10 hours, after which the cell proceeds with DNA replication (S phase). How does the expression level of HMGB1 and other proteins influence the cell's fate? We varied the levels of HMGB1 and AKT to determine how they affect cell behavior. A number of studies have found that HMGB1 is overexpressed in many cancers, and the overexpression of HMGB1 and its receptors can promote cancer cell proliferation and decrease apoptosis [8, 9]. In Fig. 4A-B, we increase the initial values of HMGB1 from 1 to 106 and measure p53's maximum expression level in phase G1. We then measure the oncoproteins E2F and CyclinD/E's expression levels at 10 hours, which corresponds to the G1-S phase transition point. For the stochastic simulation, the experiment is repeated 10 times per value to compute the mean and standard errors. Fig. 4(A, D) demonstrates that the increase of HMGB1's initial value will lead to a decrease of p53's expression level, but when the number of HMGB1 molecules is over 105, p53 will not continue to decrease. This is because HMGB1 can also activate and increase the expression level of its downstream protein E2F (Fig. 4(A, D)), whose overexpression will activate the transcription of the tumor suppressor protein ARF, which can inhibit MDM2's activity to stabilize p53's level. However, ARF is found to be mutated in up to 80% of pancreatic cancers [36, 53]. This means that ARF cannot inhibit the activity of the oncoprotein MDM2, thereby leading to lower levels of the tumor suppressor p53.

Figure 4
figure 4

HMGB1 and AKT sensitivity study. Overexpression of HMGB1 (A-B, D-E) leads to the increase of oncoprotein E2F and DNA replication proteins CyclinD/E, and to the decrease of p53. Overexpression of AKT (C, F) activated by HMGB1 increases the expression level of MDM2 p and represses p53, using ODE (A-C) and SSA (D-F) simulations.

Fig. 4(B, E) shows that the cell cycle regulatory proteins CyclinD/E will increase with the elevated expression of HMGB1, a behavior which could be verified by future experiments. Fig. 4(A-B, D-E) explains the experimental discovery that the overexpression of HMGB1 decreases apoptosis and promotes DNA replication and proliferation in cancer cells.

The oncoprotein AKT is overexpressed in many cancers [54]. In Fig. 4(C, F), we first increase the number of unphosphorylated AKT molecules and fix the other proteins' concentration, then measure p53 and MDM2 p 's expression levels at 10 hours in phase G1 after HMGB1 activates its receptor RAGE. Fig. 4(C, F) shows that with the increase of AKT's expression level, p53 is repressed due to the ubiquitination initiated by the overexpressed MDM2 p , which is promoted by the activated and overexpressed AKT protein. The results in Fig. 4(C, F) suggest a way to inhibit tumor cell proliferation and induce tumor cell apoptosis through the inhibition of protein phosporylation events downstream from AKT kinases in the PI3K/AKT pathway, using an AKT kinase inhibitor (such as the drug GSK-690693 [55]).

K-RAS is a member of the RAS protein family. K-RAS mutation and ARF loss occur in more than 80% of pancreatic cancers [36, 53]. The P21 and FBXW7 proteins are also frequently mutated in many cancers [37]. ARF and P21 play an important role in the crosstalk between the p53 and Rb pathways. ARF is able to reroute cells with oncogenic damage to p53-dependent fates through binding to MDM2 and targeting its degradation. The p53-dependent tumor suppressor proteins P21 and FBXW7 can inhibit CyclinD/E's activity to prevent the proliferation of cancer cells.

Fig. 5 shows how mutations of ARF, P21 and FBXW7, and K-RAS influence tumor suppressor and cell cycle regulatory protein levels at 10 hours in the HMGB1 signaling pathway. We use the MDM2 degradation rate driven by ARF, d ARF ( d ′ 7 in the ODE model), to describe ARF mutations. Also, we use the Cyclin degradation rate driven by P21 (dP 21for stochastic simulation, and b ′ 6 for ODE simulation) to describe P21 and FBXW7 mutations. Large d ARF and dP 21values correspond to small mutations of ARF and P21 respectively, while small d ARF and dP 21values correspond to large ARF and P21 mutations in the cell.

Figure 5
figure 5

Mutations of ARF, P21 and RAS affect the cell's fate. Mutations of the tumor suppressor proteins ARF and P21, and of the oncoprotein RAS affect the cell's fate, using ODE (A-C) and stochastic (D-F) simulations. The mutations of ARF (A, D) and P21 (B, E), and RAS (C, F), which correspond to small d ARF and dP 21, and large d RAS values, upregulate the expression level of the oncoproteins MDM2 and CyclinD/E, and downregulate p53's expression level.

Fig. 5(A, D) shows that wild-type ARF (large d ARF ) can decrease the number of MDM2 p molecules and increase p53's expression level to initiate apoptosis even if the cell proceeds to the S phase. Moreover, mutated ARF (smaller d ARF ) can not stabilize p53 expression and prevent the proliferation of cancer cells if HMGB1 is overexpressed. This could explain the phenomenon that ARF loss exists in over 80% of pancreatic cancers [36]. Fig. 5(B, E) demonstrates that CyclinD/E proteins will increase if P21 is mutated (smaller dP 21), thereby accelerating cell cycle progression.

K-RAS is mutated in most cancers, especially in pancreatic cancer [31]. The activation of RAS is initiated by HMGB1 and its receptors, and the wild-type RAS can be deactivated by some kinases. Studies have found that the mutated K-RAS can not be deactivated [56], even if HMGB1 is knocked out, so it will continuously activate the downstream signaling pathways which promote cell proliferation. Fig. 5(C, F) shows that with the increase of RAS deactivation rate d RAS (b1 in the ODE model), the synthesis of CyclinD/E will be inhibited, but a small deactivation rate of RAS will lead to overexpression of CyclinD/E. The results visualized in Fig. 5 suggest some ways to inhibit cancer cell proliferation through inhibition or deactivation of the signaling pathways involving RAS, Cyclin, and Cyclin-dependent kinases (CDK). Recently, CDK and RAS inhibitor drugs [57–59] have been developed to inhibit tumor growth.

Verification of the HMGB1 model

We use Statistical Model Checking (SMC) to verify some fundamental properties that our model should satisfy. We test whether the model satisfies a given BLTL property with probability p ≥ 0.9. We set the threshold T = 1000 for the verification, so the probability of a wrong answer is smaller than 10-3.

Property 1: p53 is normally expressed at low levels in human cells. We verified the following property

P r ≥ 0.9 [ F t ( G 900 ( P 53 < 3.3 × 10 4 ) ) ] ,

which informally means that the number of p53 molecules will be less than a threshold value within t minutes, and it will always stay below this value during the next 900 minutes. We verified this property with various values of t and the results are shown in Table 2.

Table 2 Verification of property 1

Property 2: p53's expression level increases quickly in response to various stresses, including the activation of HMGB1. We verified the property

P r ≥ 0.9 [ F 100 ( P 53 > 5.3 × 10 4 ) ] ,

that is, within 100 minutes p53's level will eventually be larger than 5.3 × 104. SMC accepts this property as true, after sampling 38 traces (of which 37 satisfying traces).

Property 3: PI3K will be activated within a few minutes after HMGB1 binds to RAGE. We verified the following property

P r ≥ 0.9 [ F 20 ( P I 3 K a / P I 3 K t o t > 0.5 ) ] ,

which means that half of PI3K will be activated within 20 minutes. We verified this property with various values of HMGB1, and the results are shown in Table 3. If HMGB1 was overexpressed (105), this property was accepted as true (22 satisfying traces). But if the expression level of HMGB1 was very low, the property was rejected.

Table 3 Verification of property 3

Property 4: The overexpression of HMGB1 will promote the oncoprotein CyclinE's expression before the G1-S phase transition point, thereby facilitating the G1-S phase transition. We verified the property

P r ≥ 0.9 [ F 600 ( C y c l i n E > 900 ) ] ,

that is, the number of CyclinE molecules will eventually exceed 900 within 600 minutes (10 hours). We verified this property with various values of HMGB1 and the results are shown in Table 4.

Table 4 Verification of property 4 and 5

Property 5: Mutation in K-RAS leads to continuous activation of downstream pathways and overexpression of CyclinD in the G1 phase during HMGB1-activated signaling transduction. We verified the property:

P r ≥ 0.9 [ F 600 ( C y c l i n D > 900 ) ] ,

with different RAS deactivation rates (d RAS ). The results are presented in Table 4. Properties 4 and 5 show that the overexpression of HMGB1 and mutation of RAS (small d RAS value) will accelerate the expression of cell regulatory protein CyclinD/E to promote cell proliferation. However, inhibition of HMGB1 and an increase of RAS deactivation rate will prevent tumor growth.

Property 6: Within 300 minutes, CyclinE's expression level becomes very low until 50% of RAS has been activated by HMGB1. We verified the property:

P r ≥ 0.9 [ ( C y c l i n E < 10 ) U 300 ( R A S a / R A S t o t > 0.5 ) ] .

SMC accepted this property as true (22 satisfying traces).

Property 7: HMGB1 could influence the tumor suppressor protein p53's expression level, especially the first peak of p53's concentration in the G1 phase. We verified the following property:

P r ≥ 0.9 [ F 100 ( p 53 10000 ≥ a ∧ F 100 ( p 53 10000 ≤ 0.4 ) ) ] ,

which informally means that the number of p53 protein molecules in the nucleus will eventually be greater than a threshold value a × 104 within 100 minutes, after which it reduces to a low level within the next 100 minutes. We verified this property with various values of a and HMGB1, and the results are shown in Table 5.

Table 5 Verification of property 7

Conclusions

We have presented a reaction network model of the signaling transduction initiated by HMGB1. The model incorporates the contributions from the most important known signaling components of the HMGB1 signal transduction network. The model is expressed in the form of BioNetGen rules, and simulated using ODEs and Gillespie's algorithm under a range of conditions. We used Statistical Model Checking to automatically validate our model with respect to known experimental results.

Our simulations demonstrate a dose-dependent p53 and CyclinE response curve to increasing HMGB1 stimulus. This hypothesis could be tested by future experiments. In particular, overexpression of HMGB1 promotes the cell cycle regulatory proteins E2F and CyclinD/E and inhibits the pro-apoptotic protein p53, leading to increased cancer cell survival and decreased apoptosis. This is consistent with experimental observations in recent studies of cancer cells [11]. We also investigated the roles of different components in the pathway and predicted their activity in response to various conditions. We investigated how mutations of the RAS, ARF and P21 proteins influence the fate of the cancer cell. In particular, parameter variation showed that the mutated RAS increases the expression level of CyclinE, leading to cancer cell proliferation. Mutation or loss of the ARF protein leads to high MDM2 activity and loss of p53 expression in the face of HMGB1 overexpression, resulting in decreased apoptosis. Our model shows that the inhibition (or deactivation) of RAS, Cyclin, and Cyclin-dependent kinases (CDK) might inhibit tumor growth.

Since our proposed model is based on just three signaling pathways, we are far from capturing the entire HMGB1 network dynamics. Studies have found that HMGB1 can not only activate the PI3K-AKT and RAS-ERK pathways, but can also activate the NFκ B signaling pathway [27], which regulates many pro-apoptotic and anti-apoptotic proteins' transcription [60]. Since HMGB1 could be released passively during necrosis, there might exist crosstalk between the tumor necrosis factor (TNF) pathway and the HMGB1 pathway. Besides the incorporation of new pathways, recent work has demonstrated that HMGB1 can bind to p53 directly to influence p53-mediated transcriptional activity [61]. A larger network for HMGB1 signal transduction will be explored in our future work.

It has been recently observed that pancreatic tumor cells increase autophagy [11] and release HMGB1 [10] in response to chemotherapy, radiation, and hypoxia, which may promote tumor cell survival. It has been hypothesized that direct inhibition of autophagy may be another way to inhibit tumor growth and enhance the efficacy of cancer therapies [11]. The incorporation of autophagic proteins into the HMGB1 signaling pathway is worth considering in future work.

Although our current model can only qualitatively compare with the experimental behavior, it still provides valuable information about the behavior of HMGB1 signal transduction in response to different stimuli. Future experiments will enable the development of more realistic models. We anticipate that the application of model checking techniques, such as those explored in this work, will facilitate the development of targeted and effective anti-cancer therapies.

References

  1. Hanahan D, Weinberg RA: The hallmarks of cancer. Cell 2000, 100: 57–70. 10.1016/S0092-8674(00)81683-9

    Article  CAS  PubMed  Google Scholar 

  2. Scaffidi P, Misteli T, Bianchi M: Release of chromatin protein HMGB1 by necrotic cells triggers inflammation. Nature 2002, 418: 191–195. 10.1038/nature00858

    Article  CAS  PubMed  Google Scholar 

  3. Bonaldi T, Talamo F, Scaffidi P, Ferrera D, Porto A, Bachi A, Rubartelli A: Monocytic cells hyperacetylate chromatin protein HMGB1 to redirect it towards secretion. EMBO J 2003, 22: 5551–5560. 10.1093/emboj/cdg516

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  4. Wang H, Bloom O, Zhang M, Vishnubhakat J, Ombrellino M, Frazier A, Yang H: HMG-1 as a late mediator of endotoxin lethality in mice. Science 1999, 285: 248. 10.1126/science.285.5425.248

    Article  CAS  PubMed  Google Scholar 

  5. Dumitriu IE, Baruah P, Valentinis B, et al.: Release of high mobility group box 1 by dendritic cells controls T cell activation via the receptor for advanced glycation end products. The Journal of Immunology 2005, 174: 7506–7515.

    Article  CAS  PubMed  Google Scholar 

  6. Semino C, Angelini G, Poggi A, Rubartelli A: NK/iDC interaction results in IL-18 secretion by DCs at the synaptic cleft followed by NK cell activation and release of the DC maturation factor HMGB1. Blood 2005, 106: 609–616. 10.1182/blood-2004-10-3906

    Article  CAS  PubMed  Google Scholar 

  7. Sparvero L, Asafu-Adje D, Kang R, Tang D, Amin N, et al.: RAGE (Receptor for Advanced Glycation Endproducts), RAGE ligands, and their role in cancer and inflammation. Journal of Translational Medicine 2009, 7: 17. 10.1186/1479-5876-7-17

    Article  PubMed Central  PubMed  Google Scholar 

  8. Ellerman JE, Brown CK, de Vera M, Zeh HJ, Billiar T, et al.: Masquerader: high mobility group box-1 and cancer. Clinical Cancer Research 2007, 13: 2836–2848. 10.1158/1078-0432.CCR-06-1953

    Article  CAS  PubMed  Google Scholar 

  9. Lotze MT, Tracey K: High-mobility group box 1 protein (HMGB1): nuclear weapon in the immune arsenal. Nature Reviews Immunology 2005, 5: 331–342. 10.1038/nri1594

    Article  CAS  PubMed  Google Scholar 

  10. Vakkila J, Lotze MT: Inflammation and necrosis promote tumour growth. Nature Reviews Immunology 2004, 4: 641–648. 10.1038/nri1415

    Article  CAS  PubMed  Google Scholar 

  11. Kang R, Tang D, Schapiro NE, Livesey KM, Farkas A, Loughran P, Lotze MT, Zeh HJ: The receptor for advanced glycation end products (RAGE) sustains autophagy and limits apoptosis, promoting pancreatic tumor cell survival. Cell Death Differ 2010, 17(4):666–676. 10.1038/cdd.2009.149

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  12. Brezniceanu ML, Volp K, Bosser S, Solbach C, Lichter P, et al.: HMGB1 inhibits cell death in yeast and mammalian cells and is abundantly expressed in human breast carcinoma. FASEB Journal 2003, 17: 1295–1297.

    CAS  PubMed  Google Scholar 

  13. Vogelstein B, Lane D, Levine AJ: Surfing the p53 network. Nature 2000, 408: 307–310. 10.1038/35042675

    Article  CAS  PubMed  Google Scholar 

  14. Minella AC, Grim JE, Welcker M, Clurman BE: p53 and SCFFbw7 cooperatively restrain cyclin E-associated genome instability. Oncogene 2007, 26: 6948–6953. 10.1038/sj.onc.1210518

    Article  CAS  PubMed  Google Scholar 

  15. Pestov DG, Strezoska A, Lau LF: Evidence of p53-dependent cross-talk between ribosome biogenesis and the cell cycle: effects of nucleolar protein Bop1 on G1/S transition. Molecular and Cellular Biology 2001, 21(13):4246–4255. 10.1128/MCB.21.13.4246-4255.2001

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  16. Wee KB, Aguda BD: Akt versus p53 in a network of oncogenes and tumor suppressor genes regulating cell survival and death. Biophysical Journal 2006, 91: 857–865. 10.1529/biophysj.105.077693

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  17. Wee KB, Surana U, Aguda BD: Oscillations of the p53-Akt Network: Implications on Cell Survival and Death. PloS One 2009, 4: e4407. 10.1371/journal.pone.0004407

    Article  PubMed Central  PubMed  Google Scholar 

  18. Ciliberto A, Novak B, Tyson J: Steady States and Oscillations in the p53/Mdm2 Network. Cell Cycle 2005, 4(3):488–493.

    Article  CAS  PubMed  Google Scholar 

  19. Zhang T, Brazhnik P, Tyson J: Exploring mechanisms of the DNA-damage response: p53 pulses and their possible relevance to apoptosis. Cell Cycle 2007, 6: 85-e10. 10.4161/cc.6.16.4577

    Article  PubMed  Google Scholar 

  20. Puszynski K, Hat B, Lipniacki T: Oscillations and bistability in the stochastic model of p53 regulation. Journal of Theoretical Biology 2008, 254: 452–465. 10.1016/j.jtbi.2008.05.039

    Article  CAS  PubMed  Google Scholar 

  21. Bottani S, Grammaticos B: Analysis of a minimal model for p53 oscillations. Journal of Theoretical Biology 2007, 249: 235–245. 10.1016/j.jtbi.2007.04.026

    Article  CAS  PubMed  Google Scholar 

  22. Clarke EM, Grumberg O, Peled DA: Model Checking. MIT Press; 1999.

    Google Scholar 

  23. Jha SK, Clarke EM, Langmead CJ, Legay A, Platzer A, Zuliani P: A bayesian approach to model checking biological system. CMSB, Volume 5688 of LNCS 2009, 218–234.

    Google Scholar 

  24. Langmead CJ: Generalized Queries and Bayesian Statistical Model Checking in Dynamic Bayesian Networks: Application to Personalized Medicine. CSB 2009, 201–212.

    Google Scholar 

  25. Rizk A, Batt G, Fages F, Soliman S: On a continuous degree of satisfaction of temporal logic formulae with applications to systems biology. CMSB, Volume 5307 of LNCS 2008, 251–268.

    Google Scholar 

  26. Larris S, Levine AJ: The p53 pathway: positive and negative feedback loops. Oncogene 2005, 24: 2899–2908. 10.1038/sj.onc.1208615

    Article  Google Scholar 

  27. van Beijnum JR, Buurman WA, Griffioen AW: Convergence and amplification of toll-like receptor (TLR) and receptor for advanced glycation end products (RAGE) signaling pathways via high mobility group B1. Angiogenesis 2008, 11: 91–99. 10.1007/s10456-008-9093-5

    Article  PubMed  Google Scholar 

  28. Barak Y, Juven T, Haffiner R, Oren M: mdm2 expression is induced by wild type p53 activity. EMBO J 1993, 12: 461–468.

    PubMed Central  CAS  PubMed  Google Scholar 

  29. Haupt Y, Maya R, Kasaz A, Oren M: Mdm2 promotes the rapid degradation of p53. Nature 1997, 387: 296–299. 10.1038/387296a0

    Article  CAS  PubMed  Google Scholar 

  30. Blanco-Aparicio C, Renner O, Leal J, Carnero A: PTEN, more than the AKT pathway. Carcinogenesis 2007, 28: 1379–1386. 10.1093/carcin/bgm052

    Article  CAS  PubMed  Google Scholar 

  31. Downward J: Targeting RAS signalling pathways in cancer therapy. Nature Reviews Cancer 2003, 3: 11–22. 10.1038/nrc969

    Article  CAS  PubMed  Google Scholar 

  32. Yao G, Lee TJ, Mori S, Nevins J, You L: A bistable Rb-E2F switch underlies the restriction point. Nature Cell Biology 2008, 10: 476–482. 10.1038/ncb1711

    Article  CAS  PubMed  Google Scholar 

  33. Sherr CJ, McCormick F: The RB and p53 pathways in cancer. Cancer Cell 2002, 2: 103–112. 10.1016/S1535-6108(02)00102-2

    Article  CAS  PubMed  Google Scholar 

  34. Sears R, Nevins J: Signaling networks that link cell proliferation and cell fate. The Journal of Biological Chemistry 2002, 277: 11617–11620. 10.1074/jbc.R100063200

    Article  CAS  PubMed  Google Scholar 

  35. Nevins JR: The Rb/E2F pathway and cancer. Human Molecular Genetics 2001, 10: 699–703. 10.1093/hmg/10.7.699

    Article  CAS  PubMed  Google Scholar 

  36. Bardeesy N, DePinho RA: Pancreatic cancer biology and genetics. Nature Reviews Cancer 2002, 2(12):897–909. 10.1038/nrc949

    Article  CAS  PubMed  Google Scholar 

  37. Mao J, Perez-losada J, Wu D, DelRosario R, Tsunematsu R, Nakayama KI, Brown K, Bryson S, Balmain A: Fbxw7/Cdc4 is a p53-dependent, haploinsufficient tumour suppressor gene. Nature 2004, 432: 775–779. 10.1038/nature03155

    Article  CAS  PubMed  Google Scholar 

  38. Hlavacek WS, Faeder JR, Blinov ML, Posner RG, Hucka M, Fontana W: Rules for modeling signal-transduction system. Sci STKE 2006, 2006(344):re6. re6 re6 10.1126/stke.3442006re6

    PubMed  Google Scholar 

  39. Bolouri H: Computational Modeling of Gene Regulatory Networks: a primer. Imperial College Press; 2008.

    Book  Google Scholar 

  40. Yang H, Hsu C, Hwang M: An Analytical Rate Expression for the Kinetics of Gene Transcription Mediated by Dimeric Transcription Factors. Journal of Biochemistry 2007, 142: 135–144. 10.1093/jb/mvm151

    Article  CAS  PubMed  Google Scholar 

  41. Alon U: An introduction to systems biology: design principles of biological circuits. Chapman & Hall; 2007.

    Google Scholar 

  42. Gillespie DT: A general method for numerically simulating the stochastic time evolution of coupled chemical reactions. Journal of Computational Physics 1976, 22(4):403–434. 10.1016/0021-9991(76)90041-3

    Article  CAS  Google Scholar 

  43. Gong H, Sengupta H, Linstedt A, Schwartz R: Simulated De Novo Assembly of Golgi Compartments by Selective Cargo Capture during Vesicle Budding and Targeted Vesicle Fusion. Biophysical Journal 2008, 95: 1674–1688. 10.1529/biophysj.107.127498

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  44. Gong H, Guo Y, Linstedt A, Schwartz R: Discrete, continuous, and stochastic models of protein sorting in the Golgi apparatus. Phys Rev E Stat Nonlin Soft Matter Phys 2010, 81: 011914.

    Article  PubMed  Google Scholar 

  45. Lipniacki T, Hat T, Faeder JR, Hlavacek WS: Stochastic effects and bistability in T cell receptor signaling. Journal of Theoretical Biology 2008, 254: 110–122. 10.1016/j.jtbi.2008.05.001

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  46. HMGB1 BioNetGen Code[http://www.cs.cmu.edu/~haijung/research/HMGB1model.bngl]

  47. Clarke EM, Emerson EA, Sifakis J: Model checking: algorithmic verification and debugging. Commun ACM 2009, 52(11):74–84. 10.1145/1592761.1592781

    Article  Google Scholar 

  48. Pnueli A: The Temporal Logic of Programs. FOCS 1977, 46–57.

    Google Scholar 

  49. Zuliani P, Platzer A, Clarke EM: Bayesian statistical model checking with application to Simulink/Stateflow verification. HSCC 2010, 243–252. full_text

    Chapter  Google Scholar 

  50. Younes HLS, Simmons RG: Probabilistic Verification of Discrete Event Systems Using Acceptance Sampling. CAV, Volume 2404 of LNCS 2002, 223–235.

    Google Scholar 

  51. Younes HLS, Simmons RG: Statistical probabilistic model checking with a focus on time-bounded properties. Information and Computation 2006, 204(9):1368–1409. 10.1016/j.ic.2006.05.002

    Article  Google Scholar 

  52. Geva-Zatorsky N, Rosenfeld N, Itzkovitz S, Milo R, Sigal A, Dekel E, Yarnitzky T, Liron Y, Polak P, Lahav G, Alon U: Oscillations and variability in the p53 system. Molecular Systems Biology 2006, 2: 2006.0033. 10.1038/msb4100068

    Article  PubMed Central  PubMed  Google Scholar 

  53. Hahn WC, Weinberg RA: Modelling the molecular circuitry of cance. Nature Reviews Cancer 2002, 2: 331–341. 10.1038/nrc795

    Article  CAS  PubMed  Google Scholar 

  54. Altomare D, Wang H, Skele K, Rienzo AD, Klein-Szanto A, Godwin A, Testa J: AKT and mTOR phosphorylation is frequently detected in ovarian cancer and can be targeted to disrupt ovarian tumor cell growth. Oncogene 2004, 23: 5853–7. 10.1038/sj.onc.1207721

    Article  CAS  PubMed  Google Scholar 

  55. Rhodes N, Heerding D, Duckett D, et al.: Characterization of an Akt kinase inhibitor with potent pharmacodynamic and antitumor activity. Cancer Research 2008, 68: 2366. 10.1158/0008-5472.CAN-07-5783

    Article  CAS  PubMed  Google Scholar 

  56. van Krieken J, et al.: KRAS mutation testing for predicting response to anti-EGFR therapy for colorectal carcinoma: proposal for an European quality assurance program. Virchows Arch 2008, 453: 417–431. 10.1007/s00428-008-0665-y

    Article  CAS  PubMed  Google Scholar 

  57. McInnes C: Progress in the evaluation of CDK inhibitors as anti-tumor agents. Drug Discovery Today 2008, 13(1920):875–881. 10.1016/j.drudis.2008.06.012

    Article  CAS  PubMed  Google Scholar 

  58. Malumbres M, Pevarello P, Barbacid M, Bischoff J: CDK inhibitors in cancer therapy: what is next? Trends in Pharmacological Sciences 2008, 29: 16. 10.1016/j.tips.2007.10.012

    Article  CAS  PubMed  Google Scholar 

  59. Rotblat B, Ehrlich M, Haklai R, Kloog Y: The Ras inhibitor farnesylthiosalicylic acid (Salirasib) disrupts the spatiotemporal localization of active Ras: a potential treatment for cancer. Methods in Enzymology 2008, 439: 467–489. full_text

    Article  CAS  PubMed  Google Scholar 

  60. Lipniacki T, Paszek P, Brasier A, Luxon B, Kimmel M: Crosstalk between p53 and nuclear factor-kB systems: pro-and anti-apoptotic functions of NF-kB. Journal of Theoretical Biology 2004, 228: 195–215. 10.1016/j.jtbi.2004.01.001

    Article  CAS  PubMed  Google Scholar 

  61. Imamura T, Izumi H, Nagatani G, Ise T, Nomoto M, Iwamoto Y, Kohno K: Interaction with p53 enhances binding of cisplatin-modified DNA by high mobility group 1 protein. J Biol Chem 2001, 276: 7534–7540. 10.1074/jbc.M008143200

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This work was supported by a grant from the U.S. National Science Foundation's Expeditions in Computing Program (award ID 0926181). The authors thank Michael T. Lotze (University of Pittsburgh) for calling their attention to HMGB1 and for helpful discussions of the topic. H.G. would like to thank Marco E. Bianchi (San Raffaele University) for email discussions on HMGB1. We would also like to thank Ilya Korsunsky and Màtè L. Nagy for their comments on this paper.

This article has been published as part of BMC Bioinformatics Volume 11 Supplement 7, 2010: Ninth International Conference on Bioinformatics (InCoB2010): Bioinformatics. The full contents of the supplement are available online at http://www.biomedcentral.com/1471-2105/11?issue=S7.

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Haijun Gong or Paolo Zuliani.

Additional information

Competing interests

The authors declare that there are no competing interests.

Authors' contributions

E.M.C., J.R.F., H.G. and P.Z. proposed the project; H.G. and P.Z. wrote the manuscript; H.G. wrote the BioNetGen code and performed the numerical simulations and formal verifications; P.Z. wrote the statistical model checker; A.K. wrote the model checker code for BioNetGen. All authors read and approved the final manuscript.

Electronic supplementary material

12859_2010_4248_MOESM1_ESM.pdf

Additional file 1: Ordinary differential equations and model parameters. The PDF file contains all the ordinary differential equations that describe the HMGB1 signal transduction model, the input parameters and their descriptions. (PDF 121 KB)

Rights and permissions

This article is published under license to BioMed Central Ltd. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Gong, H., Zuliani, P., Komuravelli, A. et al. Analysis and verification of the HMGB1 signaling pathway. BMC Bioinformatics 11 (Suppl 7), S10 (2010). https://doi.org/10.1186/1471-2105-11-S7-S10

Download citation

  • Published:

  • DOI: https://doi.org/10.1186/1471-2105-11-S7-S10

Keywords